Chapter 8 - Emulsions

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 94

8 EMULSIONS

PIETER WALSTRA

Emulsions can be formed and used in various situations:


- Several organisms produce emulsions, the prime example being the milk excreted
by female mammals. Milk contains small oil droplets dispersed in an aqueous liquid.
The function of the droplets is to transport substances that do not, or insufficiently,
dissolve in water: this includes the transport of a large quantity of edible energy in a
limited volume without increasing the osmotic pressure.
- Most types of emulsions encountered in daily life are man-made. Their most
common function is the transport of water-insoluble substances in a stable, and hence
finely dispersed form, but other, more specific functions are also involved. Such
products include a range of foodstuffs, pharmaceuticals, cosmetics, pesticide formul-
ations, paints, lubricants, and finishing agents.
- Emulsions can be used in intermediate stages in manufacturing processes. This
may concern extraction in a stirred tank or a column. Another example is the form-
ation of latices by emulsion polymerization ^^
- Emulsions can also be a nuisance, since they can be formed inadvertently during
some processes, and then have to be broken. An example is crude oil, which is often
obtained as a water in oil emulsion.
The fundamentals of interface and colloid science can be applied fruitfully to
understand the various properties of emulsions and to predict how these properties
can be realized. Interfacial properties and processes are essential in formation and
stability of emulsions, and the interfacial area is large. If the drops are not too large,
emulsions can be treated as lyophobic colloids. In many respects, they form ideal
systems for studying such colloids. The particles are perfect spheres and, for many
purposes, they can be considered to be rigid. Several emulsion properties, such as
volume fraction, average droplet size, and interfacial composition, can be varied as
desired. Polydispersity of the drops may pose a problem, but there are ways of
obtaining almost homodisperse emulsions, at least in some systems^^
On the other hand, it is often far from easy to understand what happens in
practice, let alone to predict what will happen. This is due to the complexity of the

^^ See e.g. the review by J.W. Vanderhoff, Chem. Eng. Set 48 (1993) 203.
^' J. Bibette, J. Colloid Interface Set 147 (1991) 474.
Fundamentals of Interface and Colloid Science, Volume V © 2005 Elsevier Ltd.
J. Lyklema (Editor) All rights reserved
8.2 EMULSIONS

phenomena involved. First, several different changes can occur simultaneously, both
during emulsification £ind at rest, and each of these may depend in a different mainner
on internal and external variables. Second, the conditions during the application and
storage of the emulsion may vary widely; for example, chemical changes can occur that
affect physical properties. Moreover, several products are not simple emulsions, but
contain other structurad elements such as gas bubbles or solid particles, or meso-
morphic phases.
In this chapter, we shall primarily consider simple o/w (oil in water) or w/o (water
in oil) emulsions, excluding microemulsions (see chapter 5) and very coarse emul-
sions. The average droplet diaimeter is typically of the order of a micrometer. Even
such a simple emulsion may contain numerous components, for example, the
emulsifiers used in practice are virtually always mixtures.

8.1 Characterization

8.1a Description
An emulsion is a dispersion of drops in another liquid where the two liquids are
not miscible in all proportions; it is thus a two phase system. The phases are, for
simplicity, called oil and water. In this context 'oil' means a hydrophobic liquid, e.g.
benzene, hexadecane, or a triglyceride oil; and 'water' means an aqueous solution.
Nearly all emulsions contain at least one other substaince, generally Ccilled an
"emulsifier". It is needed in the formation of an emulsion, and the emulsifier's
properties then adso determine whether the emulsion will be of the o/w or of the w/o
type. The emulsifier is generally also needed to stabilize the drops in the finished
emulsion against a^regation aind coalescence. All emulsifiers are surfactants that
adsorb onto oil water interfaces, lowering the interfacial tension. Since the surfactant
must be soluble in at least one of the two phases in order to be active during
emulsification, the adsorption layer formed will generally be a Gibbs type monolayer.
Emulsions differ from suspensions in that the particle surface is deformable, both
in directions perpendicular and parallel to the surface. This permits a relatively
simple way of formation of an emulsion from two bulk phases. It also causes the
particles to assume a spherical shape. On the other hand, the droplets behave like
rigid psirticles under most conditions. Consider a drop of diameter d = l | i m , and
interfacial tension y = 0.005 mN m"^. Its Laplace or capillary pressure 4:y/d (see sec.
1.2.23) will then equal 20 kPa. Assume further that the liquid around the drop is
stirred, that it has a viscosity of 0.002 Pa s, and that the local shear rate is 5x10-^ s"^,
which is a very high value. The shear stress acting on the drop then will equal 10 Pa,
much smadler than the Laplace pressure. As we will see in sec. 8.2b, sub [i], the stress
would lead to a maximum relative deformation of the droplet by 0.1%, a negligible
amount. The shear stress can, in principle, also move the droplet's surface in a
t£ingential direction, at least locally. However, the stress will induce an interfacial
EMULSIONS 8.3

tension gradient, which means that a counteracting stress will develop, which can stop
the motion. This is discussed further in sec. 8.1c.
For very small droplets, say with a radius <0.1|Lim, the tendency of many
surfactants to impose a 'spontaneous curvature' to a monolayer, will become import-
ant (see Vol. Ill, sec. 4.7). We then have a regime between that of micro emulsions and
macro emulsions. In this chapter we will primarily consider emulsion droplets that
are large enough for the spontaneous curvature of the surfactant to have no effect, but
small enough for the drop to be rigid under most conditions.
In many respects, an o/w emulsion resembles an aqueous foam. However, foam
bubbles tend to be one or two orders of magnitude larger than emulsion droplets, and
the difference in density between both phases is larger by about one order of
magnitude. This means that foam bubbles cream very rapidly, to form a packed layer.
Liquid then drains from that layer and the Laplace pressure of the bubbles is small
enough to allow their deformation. In this way a polyhedral foam is formed. Such a
transition generally does not occur in emulsions, except when they are centrifuged.

Important internal variables affecting emulsion properties include:


- Emulsion type, i.e., o/w or w/o. The macroscopic properties of an emulsion are
generally dominated by those of the continuous phase, especially If the volume
fraction of droplets (^) is not very high. It also determines with which of the two
liquids the emulsion can be diluted. More complicated types are multiple emulsions,
e.g. w/oAv and o/w/o^'. Water in water emulsions can be formed from a solution of two
incompatible polymers, but the emulsion aspects of these phase separated systems
have hardly been studied; the inter facial tension between the phases is very small.
- Surfactant type and concentration. These determine, in combination with the spe-
cific interfacial area (A^), the surface excess (/I and the surface layer composition,
and thereby the interfacial tension and the interfacial rheological properties. Moreover,
the colloidal interaction forces between droplets depend strongly on the surface layer
composition, and thereby the stability of the emulsion against aggregation and coales-
cence. Surfactant remaining in solution in the continuous phase can also affect aggreg-
ation, depending on type and concentration.
- Properties of both phases. This concerns, for example, viscosity (77), mass density
(p), dielectric constant, ionic strength, and pH. All these variables may affect physical
stability. The mutual solubility of both liquids is also important: significant solubihty
of the disperse phase in the continuous one will promote Ostwald ripening. The
composition of both phases determines the properties of the pristine interface between
them, and then also the equilibrium composition of the interfacial absorption layer,
with all its consequences.

' See e.g., E. Dickinson and D.J. McClements, Advances in Food Colloids, Blackie, (1995),
chapter 9, which includes a general discussion of the formation and the various instabilities of
multiple emulsions.
8.4 EMULSIONS

- Volume fraction of droplets. This is often taken to equal the volume fraction of the
disperse phase, but if the droplets are quite small, the adsorption layer may
contribute substantially to (p. Its value markedly affects some physical properties,
especially the rheology, of the emulsion, and it may also affect coalescence stability.
For (p ~ 0.7, all globules touch their neighbours, and for higher (p values the drops
will increasingly deform each other.
- Size distribution of the droplets. In most cases, the stability of an emulsion is
greater if the droplets are smaller, except with respect to Ostwald ripening. Hence, the
size distribution is an important variable, both the average size and the distribution
width. The volume surface average droplet diameter, d^2' ^^ rarely below 0.3 |im , and
it may be up to a few times 10 |Lim. An important variable determined by the size
distribution is the total specific surface area of the droplets, given by,

A^=6^/d32 [8.1.1]

in units of area per unit of emulsion volume.

8.1b Surfactants as emulsifiers


In this section we proceed on treatments given in chapters III.3 and 4, and IV.4.
A surfactant used as an emulsifier has two main functions: allowing emulsion
formation and providing stability to the emulsion once made. It generally also
determines the emulsion type, o/w or w/o. Emulsifiers come in two main types: low
molar mass amphiphilic compounds, called 'amphiphiles', here, for convenience
(although it may be noted that some authors reserve the words surfactant or
emulsifier for 'amphiphile'), and surface active polymers. These differ in several
respects, as discussed below. Moreover, properties such as the interfacial tension and
the surface excess obtained vary among types of interface: s-w, s-o, a-w and o-w. The
differences can be large, and for water soluble surfactants, much more is known about
the behaviour at the a w than at the o w interface. For many polymers, nearly all of the
literature concerns solid interfaces. Moreover, for a given surfactant, surface activity
and interfacial properties may vary substantially with the composition of the aqueous
phase and that of the oil phase. For example, benzene, a liquid paraffin, or a
triacylglycerol mixture, as the oil phase will result in different / values. It is generally
difficult, and often impossible, to find adequate and rehable data.

(i) Amphiphiles. The hydrophobic part of the molecule is generally an aliphatic


chain, e.g., of palmitic or oleic acid, to which a polar head group is esterified. The
head groups, and thereby the amphiphiles, are classified as non-ionic, anionic and
cationic. In principle, the adsorbed and the dissolved surfactant are in thermodynamic
equilibrium, as given by Gibbs' adsorption law (see 1.2.22). This means that the
amphiphile forms a Gibbs monolayer (see III.4). It also means that for an emulsion of
EMULSIONS 8.5

given composition, specific surface area, and temperature, the surface properties {y,
r, etc.) are fixed.
Several surfactant properties are of importance for application in emulsions. These
include:
Solubility in both phases, and also the critical micellization concentration (c.m.c),
if it exists. The solubility depends greatly on temperature. Many water soluble
amphiphiles have a so called Krafft temperature, below which they form crystals
(mostly a-crystals), and leave a very small concentration in solution. Several amphi-
philes can also form crystals in an oil phase. For water soluble amphiphiles, a plateau
value for the interfacial tension is generally reached at the c.m.c. or at the solubility
limit; these concentrations are, for the most part, between 0.01 and 10 mmolar, or
between 10 mg and 3 g per litre, the lower values generally applying to non-ionics.
Hydrophile Lipophile Balance (HLB)^\ This is a measure of the preponderance of
the hydrophilic over the lipophilic moiety of a surfactant molecule at 25°C. The HLB
number changes linearly with the molar Gibbs energy of micellization, and arbitrary
scaling constants are introduced in such a way that the minimum HLB number equals
about zero, and the number for a molecule that is neither hydrophilic nor lipophilic
about seven. Some methods exist to determine HLB numbers; they can also be
estimated by summation of numbers assigned to specific groups of the molecule.
Tabulated values are available. A surfactant with a low HLB number (2-6) is better
soluble in oil than in water and tends to make a w/o emulsion, and for higher HLB
numbers (10-18) it is the other way round. This is in agreement with Bancroft's rule
(see Vol. Ill, sec. 4.8): the continuous phase of an emulsion is the one in which the
surfactant is best soluble. There are, however, exceptions'^ The numbers given refer
to non-ionics. Most ionic amphiphiles have quite high HLB numbers and are poorly
soluble in oil.
A related concept is that of the phase-inversion temperature (p.i.t.) or HLB
temperature. It applies to non-ionic amphiphiles containing one or more poly
oxyethylene chains. The p.i.t. is defined as a property of the emulsion, since it depends
not only on the surfactant but also on the compositions of both phases. At a
temperature below the p.i.t., the surfactant tends to make an o/w emulsion. If the
temperature is increased, the solubility of the surfactant in water decreases (cf. Vol.
Ill, sec. 4.6c), XQ^ decreases to a very low value, and the emulsion becomes very
unstable to coalescence. At the p.i.t., y reaches a minimum. Upon further increase in
the temperature, the surfactant becomes increasingly soluble in the oil phase, y
increases again, and a w/o emulsion forms. For a homologous series of surfactants.

K. Shinoda and H. Kunieda, in P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 1,


chapter 5, p. 337, Marcel Dekker (1983).
B.P. Binks, in B.P. Binks, Ed., Modern Aspects of Emulsion Science, chapter 1, p. 1. Royal
Soc. Chem. (1998).
8.6 EMULSIONS

the p.i.t. increases with increasing number of oxyethylene groups. Broadly speaking,
the HLB number and p.i.t. are positively correlated.
The interfacial tension caused. In most emulsions, sufficient surfactant is present
to reach a plateau value. This is, for the most part, between 2 and 5 mN m"^ at
triglyceride oil water interfaces, and often has somewhat higher values at paraffin oil-
wat(^r interfaces. Oxyethylated surfactants can give much lower values near the p.i.t.
Some anionic surfactants give very low y values at a specific ionic strength, e.g.
< 1 |j V m-^ for Aerosol OT at 0.05 molar NaCl^^

Figure 8.1. Interfacial tension


( 7Q^ ) against temperature for
cottonseed oil/water systems
with 1 monostearin or 1 mono-
palmitin (broken curve) in the
oil at various concentrations
(indicated in %). (Redrawn from
data by button et al., loc. cit.)

For many non oxyethylated amphiphiles, y^^ decreases with decreasing temperat-
ure. For those with a saturated aliphatic chain, a sharp break in the y{T) curve can
occur, as shown in fig. 8.1, if the surface excess is high^\ The temperature at which
this occurs, T^, is called the chain crystallization temperature. It is assumed that the
aliphatic chains crystallize in the interface, as in an a-crystal of the amphiphile. The
molar melting enthalpy in the interface is of the same order as, but somewhat smaller
than, that of a-crystals. T^ is lower by 20 to 50 K than the melting temperature of the
crystal, the difference being smaller for a higher value of F. The plateau values of F
differ greatly, below and above T^, being, e.g., 6 and 3 |imol m~^ , respectively, for
monoglycerides; the respective values of y^^ are, e.g., 7 and 20 mN m"^.
The importance of a low y value is that it facilitates the break-up of droplets into
smaller ones during emulsification. On the other hand, the stability against coales-
cence tends to be smaller.
Surface layer composition. Often, the surface excess equals, again, about the
plateau value, for the most part one to a few mg m"^. The composition of the surface
layer is important for the stabilization of emulsion droplets against aggregation and
coalescence. Electrostatic repulsion can occur for ionic surfactants, and it depends

^' R. Aveyard, B.P. Binks, S. Clark, and J. Mead, J. Chem. Soc. Faraday Trans. I 82 (1986)
125.
2) E.S. button, C.E. Stauffer, J.B. Martin, and A.J. Fehl, J. Colloid Interface Sci. 30 (1969) 283.
EMULSIONS 8.7

mainly on the surfactant type, pH, and ionic strength. Steric repulsion can be provided
by surfactants that have flexible chains protruding in the continuous phase, e.g.,
Tweens for o/w emulsions. The repulsion depends strongly on the solvent quality for
the chains, and hence on temperature.
Interfacial rheology will be discussed in sec. 8.1c.

(ii) Mixtures of amphiphiles^K In practice, surfactant preparations are mix iires.


The length and the number of unsaturated bonds, or any branching of the h d r o -
phobic chain, may differ, and so may the composition of the hydrophilic head group.
Moreover, mixtures of amphiphiles with different properties are sometimes used
intentionally. A special case is the addition of a, 'cosurfactant', i.e., a weakly surface
active substance such as a medium chain length aliphatic alcohol. The application of
mixtures rather than pure amphiphiles has some important consequences.
The plateau value of the interfacial tension may be either higher or lower than for
pure amphiphiles. Presumably, this depends on the effect of composition on the max-
imum packing density in the interface. Mixtures, especially equimolar mixtures, of an
anionic and a cationic amphiphile can yield quite low / values (high surface pres-
sures). Something similar can happen for equimolar mixtures of comparable Spans
and Tweens (e.g., sorbitan monopalmitate and sorbitan polyoxyethylene monopalmit-
ate). However, some mixtures yield higher y values than each of the pure components.
The equilibrium composition of the interfacial monolayer in an emulsion will be
different from, and the interfacial tension will be higher than, that at a macroscopic
o/w interface. This is because the area to volume ratio v^ll be much larger in
emulsions. Relatively small quantities of a surfactant that gives a relatively low y value
can eventually dominate in the macroscopic surface layer, but will remain a minor
component in the layer on the emulsion droplets. Moreover, it may take a long time,
say an hour, to reach a constant y value at a macroscopic interface.
Many amphiphiles with long aliphatic chains are quite poorly soluble in water, and
possibly even in oil, at low temperature. They mostly form a-crystals upon cooling,
leaving too little surfactant in solution to make an emulsion. Mixtures, however, tend
to crystallize very slowly, if at all, and emulsions can readily be made.
The most important difference may be in the relatively high value of the dilational
modulus of the adsorption layer obtained with mixtures, as compared to pure surfact-
ants at high concentration, which will be explained in sec. 8.1c. It may be the main
reason why it is difficult to make emulsions (and especially foams) with a pure
amphiphile.

^^ Sec, for a general discussion, see E.H. Lucassen Reynders, in E.H. Lucassen Reyndcrs. Ed.,
Anionic Surfactants Vol. 11, Physical Chemistry of Surfactant Action, Marcel Dekker (1981),
chapter 1, p. 1.
8.8 EMULSIONS

(iii) Polymers. We will first consider synthetic polymers, whose surface properties
are discussed in chapters 1 and 2; see also I11.3.4i and III.3.8e, f. The simplest are
Unear homopolymers, such as poly (ethylene), but these are generally unsuitable for
emulsions. They tend to be either insoluble in the continuous phase or non adsorbing.
Hence, copolymers, usually of two kinds of monomers, are generally applied. An
example is the water soluble polyvinyl alcohol (PVA), which is actually a partially
hydrolyzed poly(vinyl acetate); the -OCOCH3 groups left, mostly between 2 and 12%,
are hydrophobic and tend to become lodged in an o w interface, the other segments
assure solubility in water. Two block or three block copolymers can be made, as well
as graft polymers (several short chains covalently bound to a long linear chain) and
these may have very good emulsion stabilizing properties, but most of the surfactants
used in practice are random copolymers. The conformation of the adsorbed polymer
then depends on molar mass (degree of polymerization), fraction of hydrophobic
groups, and especially on the distribution of these groups over the chain. The con-
formation, especially the protrusion of long flexible tails into the continuous phase, is
an important variable for the strength of the steric repulsion between emulsion drops.
In practice, several kinds of copolymers are used, for o/w and w/o emulsions. Some of
these are polyelectrolytes, which can also provide electrostatic repulsion.

(a) s/^ (b) J


20
2: / ^
e /c Is /c
10
p ^ ^

0 ^^1——T
0.1 10
1
100
y ^^—1 1
2
-eq/
r / m g m~^

Figure 8.2. Equilibrium values of the surface pressure, TT, as a function of concentration in the
continuous phase, c^^ , and of the surface excess, F, for the surfactants sodium dodecyl
sulfate (S), a PVA sample (P), and p- casein (C; molar mass 24 kDa). Approximate results at o-w
interfaces, meant to illustrate trends.

Figure 8.2. shows an important difference between a typical adsorbing polymer


(PVA) and most other surfactants: the concentration range over which the surface
pressure increases from zero to a plateau value is very wide, e.g., a factor of 10^
rather than 10-100. This is because of the wide variation in molar mass and in the
distribution of acetate groups. Other things being equal, the largest molecules are the
most surface active ones. This means that extensive exchange of molecules at the
surface occurs, which is a slow process because an adsorbed polymer molecule will
EMULSIONS 8.9

generally be adsorbed with many segments, and all of these must be desorbed sim-
ultaneously. Consequently, it can take a very long time, e.g., a day, before equilibrium
values of surface pressure, composition, and conformation of the monolayer are
obtained. Moreover, the relationship between surface pressure and surface excess will
depend on the polymer concentration and specific surface area, since these factors
also affect the various concentrations in the continuous phase.
Another characteristic of polymeric surfactants is that the maximum attainable
surface pressure is generally lower than that reached by most amphiphiles. Hence,
very small values of y^^ cannot be reached, although there is considerable variation
among polymers.
Nearly all natural polymers that are used as surfactants are proteins; see chapter 3.
Because they are not soluble in oils, they can only be used to make o/w emulsions. An
important difference between a pure protein and a synthetic polymer preparation is
that all the protein molecules are, in principle, identical. Some of the side groups are
hydrophobic, others are negatively or positively charged. Virtually all water-soluble
proteins adsorb onto o-w interfaces. They are very surface active, as illustrated in fig.
8.2a: the concentration at which a protein starts to significantly increase the surface
pressure is always very small, as compared to most amphiphiles. By and large,
proteins of larger molar mass are more surface active. The high surface activity is due
to the large reduction in Gibbs energy per molecule upon adsorption, e.g., 1 0 ^ - 1 0 ^
times fcT. This also means that desorption will be a very slow process. 'Washing away'
of a protein from the emulsion drops by dilution with water (or buffer) is virtually
impossible, the more so because the very high surface activity precludes the formation
of a significant concentration difference. Desorption is even slow upon compression of
a monolayer, especially for large molecules ^^
On the other hand, the plateau value of y^^ is not very small, of the order of
10 mN m"^. Figure 8.2b shows two equations of state, and it is seen that at low F
values (expressed in units of mass per unit surface area) the value of n is much
smaller for the protein than for the amphiphile. This is, for the most part, due to the
very small value of F (expressed in moles per unit surface area for proteins), taking
into account the fact that n = RTF for small F.
Proteins as emulsifiers are conveniently divided into two classes. One of these
contains proteins that form more or less random coils in solution. These include the
caseins and the gelatins (although gelatin is an exception in being not very surface
active: it has only a small proportion of hydrophobic side groups). They adsorb onto
an o w interface in a manner comparable to synthetic copolymers, with trains, loops
and tails. However, the conformation in the adsorbed state is much more fixed, since
it depends on the invariant primary structure of the protein. The conformation of
adsorbed p-casein is well known. The time needed to obtain an equilibrium

^^ F. MacRitchie, J. Coll Interface Set 105 (1985) 119.


8.10 EMULSIONS

conformation is, e.g., 10 s (at the a w interface)^\


Most of the other proteins used are globular. Upon adsorption they change con-
fornicHion, but they generally do not unfold to a considerable extent: the molecular
diami iiers in the interface and in solution tend not to be greatly different, and the
adsc bed proteins may be considered as very small deformable particles'^\ The time
scales involved in obtaining an equilibrium conformation vary between 2 and 15 min-
utes, but some of the conformational changes occur much faster. Many proteins tend
to give roughly the same relation between / and concentration at a given interface, but
the equation of state and the surface rheological properties can vary considerably^^
It may finally be noted that proteins are not always applied in their native state.
Globular proteins that have been heat denatured before adsorption show markedly
altered surface properties. The same holds for proteins that have formed aggregates,
in which case the surface coat of the droplets is not a monolayer. The surface excess
and structure of the surface layer can also depend on its history; for example, the
emulsification process itself may cause protein denaturation and a^regation.

(iv) Mixtures of polymers and amphiphiles. This combination can give rise to com-
plicated phenomena, depending on the type and the concentration of both com-
pounds^\ This applies to synthetic polymers as well as proteins.
For non-ionic amphiphiles, the situation is often relatively simple. They give a lower
Y value than most polymers (cf. fig. 8.2a) and upon increasing amphiphile concentra-
tion, they gradually displace the polymer from the interface. At first, two dimensional
phase separation occurs In the interface, and at about the c.m.c. of the amphiphile all
polymer is in solution. The rate of desorption tends to be much faster than will occur
upon compression of the interface in the absence of amphiphile.
Ionic amphiphiles often bind to polymers, even if the latter bear no charges. This
means that the activity of the amphiphile in solution is decreased owing to the
presence of polymer. This can result in a higher y value at the same total amphiphile
concentration, and the c.m.c. of the latter is apparently increased. If the polymer is
surface active, the polymer amphiphile complex will also adsorb, giving its own /
value. Nevertheless, at high amphiphile concentration, the polymer is likely to be com-
pletely displaced from the o-w interface.

H.K.A.I, van Kalsbeek and A. Prins, in E. Dickinson and J.M. Rodriguez Patino, Food Emul-
sions and Foams, Royal Soc. Chem., (1999) p. 91.
^^ J A. de Feijter and J. Benjamins, J. Coll. Interface Set 90 (1982) 289.
Several data for proteins at the interface between triacylglycerol or paraffin oil and water are
given by J. Benjamins and E.H. Lucassen Reynders, in D. Mobius and R. Miller, Eds., Proteins
at Liquid Interfaces, Elsevier (1998) p. 341.
B. Lindman, in K. Holmberg, Ed., Handbook of Applied Surface and Colloid Chemistry,
Wiley, (2002) Vol. 1, chapter 20, p. 445..
EMULSIONS 8.11

(a ) no surfactant
z
t A
oil
1 w
\\
\\
7 constant
\\
/aier ^s.
^s.
[ ^ u .

(b ) with surfactant

flow causes vy Figure 8.3. Inter facial ten-


sion gradients Vy in rela-
^ ! A U^ tion to flow of the adjacent
liquids, v = linear velocity.
Further see text. (Redrawn
from P. Walstra, Physical
Chemistry of Foods, Marcel
(c) Marangoni effect Dekker (2002).)

V 7 causes flow

8.1c The role ofinterfacial tension gradients^^


Figure 8.3a shows an interface between water and oil, devoid of surfactant, where
the water is caused to flow parallel to the interface. At the interface, the velocity
gradient Vu equals (du^/dz)Q. There is thus a tangential (shear) stress, T/^VUQ ,
acting on the interface; 77^ is the viscosity of the water phase. The interface cannot
withstand a tangential stress and the liquid velocity must thus be continuous across
the interface: the interface and oil also move. The shear stress is also continuous, but
the velocity gradient is not, because generally T]^ ^ rj^ .
In fig. 8.3b the interface contains a surfactant. The flow will now cause the
surfactant to be swept down. This implies that an interfacial tension gradient (V7) is
formed, which exerts a tangential stress of magnitude dy/dx. If the gradient can be
large enough, the stresses will be equal and opposite, hence;

This section is mainly based on Vol. Ill, sec. 3.6, especially 3.6e and 6f, and on the reviews by
Lucassen Reynders (chapter 5) and by Lucassen (chapter 6) in Lucassen Reynders (1981),
mentioned in the General References.
8.12 EMULSIONS

J ^ l =^ 18.1.2]

If so, motion of the interface and the oil phase is arrested; mechanically, the interface
acts as a solid wall for tangential stresses. However, the average Vy can never exceed
K/ doc , where Ax is the distance over which the shear stress is acting. Moreover, the
gradient can readily relax, as will be discussed below. In other words, the interface
shows viscoelastic behaviour.
In fig. 8.3c, it is seen that a /-gradient, e.g., as generated by local application of
surfactant, may cause flow on both sides of the interface. This is called the Marangoni
effect (some authors use this term also for the phenomenon depicted in fig. 8.3b); cf.
Vol. Ill, sec. 3.6e. If one of the phases has a much lower viscosity than the other (e.g.,
at an a-w surface), [8.1.2] will hold.
It is of considerable interest to know whether the o-w interface of an emulsion
droplet resists motion in the tangential direction, since it can affect droplet deform-
ation and break-up, droplet stability against coalescence, and the (bulk) rheological
properties of the emulsion. The question is whether Vy is large enough, and lasts long
enough, i.e., during the time that an external tangential stress acts on the droplet. In
fig. 8.3b the interface is subject to uniaxial dilation, and the parameter of interest then
is the (complex) surface dilational modulus, K J . In practice, the stress acting on a
droplet may involve surface shear deformation, but the dilational effect is mostly
predominant; moreover, for an interface containing amphiphiles, the surface shear
modulus tends to be far smaller than the dilational modulus.
The surface dilational modulus is given by (see Vol. Ill, sec. 3.6c, h),

K^ = dy/d\nA ( = d;r/dlnr) [8.1.3]

The equality between parentheses is only valid for a Langmuir monolayer (i.e., no
exchange of surfactants between interface and bulk), where AF is constant and,
moreover, the surfactant should not change conformation during dilation. In that case,
the modulus is purely elastic. If the deformation of the interface is uniaxial, we have
d^/dx = d ^ / x d l n A = K ^ / x , since dlnA = d l n x in this case (for isotropic biaxial
dilation din A = 2 d l n x ). Hence, the stress caused by Vy can be given by K j / x , and
for X we can use the droplet radius, a. This leads to a Marangoni number for the
ratio of the internal over the external stress, given here by

Ma = —^—= - ^ — [8.1.4]
o-ext ^Vu

where the second equality applies for an external shear stress (cf. Vol. Ill, pp. 3.97
98). For Ma » 1, the droplet surface is rigid (as in fig. 8.3b); for Ma « 1, it is fluid
(as in fig. 8.3a).
EMULSIONS 8.13

We thus need absolute values for the time-dependent dilational modulus. For a
Gibbs monolayer this is given by,

[8.1.5]
I ^ ' " ( 1 + 2^+2^)1/2

Note that the numerator equals K^ for a Langmuir monolayer. For an amphiphile
that is soluble in one of the phases and that can exchange freely between the interface
and this phase, w^hile surfactant transport is by bulk diffusion, we have.

1/2
[8.1.5a]
^ dr\2 /

where T is the characteristic time scale for deformation, e.g., l/Vu if deformation is
due to simple-shear flow. The value of ^ equals the ratio of t over the diffusion time,
2(dr/dc)2 / D . Generally, the diffusion coefficient D = O (IQ-i^ m^ s'M .

Figure 8.4. Surface dilational moduli. In (a) the dimensionless ratio Q ^\K^\/{d7i:/d\nr) is
given versus the dimensionless time, ^^: after results by J. Lucassen and M. van den Tempel,
Chem. Eng. Set 41 (1972) 1283. In (b) the value of |Kg| is given for an SDS solution of varying
concentration (c, millimolar), for a time scale of 0.02 s (after results by J. Lucassen, in
Lucassen Reynders (1981), see General References).

Some trends are illustrated in fig. 8.4. Panel (a) shows that for small ^, i.e., short
T and/or low c, the interface behaves like a Langmuir monolayer; it is as if the
surfactant is insoluble. For large ^ values, the modulus becomes quite small, the
surface properties being close to their equilibrium values. Panel (b) gives an example,
for SDS, of the dependence of the modulus on the surfactant concentration in
solution, for a given value of r . At low c the modulus rises steeply with c, in accord-
ance with the steep increase of n: with increasing F, as seen in fig. 8.2b. At high c, the
modulus tends to go to zero, because dP/dc goes to zero, and hence ^ goes to
infinity. In practice, this is generally not observed. The value of the modulus decreases
at high c values, but it remains at a significant level, because the 'surfactant' is, in fact,
8.14 EMULSIONS

a mixture of components, in different concentrations and of different surface activities.


However, at longer time scales (several minutes or more) the modulus goes to zero.
Some sample calculations may be useful. Assume that an emulsion droplet with
a = 1 |im is subject to a shear stress of 10 Pa, owing to a shear flow of Vu =
5 X 10*^ s"^; this impUes a characteristic time r of 0.2 ms. At such a small r value the
dilational modulus will probably be high, but even if it is as low a s l mNm~^, the
Marangoni number will be as high as 100. This means that the surface behaves as if it
is rigid. For a sedimenting droplet, the shear stress acting on the drop is of the order
of g aAp, where g is the acceleration due to gravity ( 1 0 m s"^ ) and Ap is the density
difference between drop and continuous phase, say 100 kgm~^. Even for a modulus
as low as 1 |LIN m~^, we still find Ma = 10*^. Also here, the interface will be perfectly
rigid. During emulsion formation the surface excess on the drop may be quite low, but
since the time scales involved are very short (e.g., 10 |is ), the modulus may still be
appreciable. These high values of Ma are, to an important extent, due to the small
value of a, hence the quite high value that Vy can have, even at a low value of ir .
However, all of the discussion above applies to amphiphiles, whereas for polymers
the trends may be quite different. Equation [8.1.5] cannot be applied, because the
polymer can, and generally will, change conformation upon expansion or compression
of the interface, which can markedly affect the values of TU . Nevertheless, the values of
the surface dilational modulus are generally high. For several proteins, at a range of
concentrations, values for \K^\ ranging between 20 and 70 mN m~^ were observed at
a time scale of about 1 s, and for some PVA samples between 10 and 15 mN m~^ .
This means, again, that the interfaces tend to be quite rigid, for the most part owing to
the slowness of any change in surface excess. In other words, the interface behaves
more like a Langmuir layer. At a longer time scale, say of an hour, the moduli tend to
be smaller, e.g., 5 mN m~^.
At short time scales, the dilational modulus would thus be given roughly by
d ; r / d l n r . Figure 8.2b shows that for p-casein \K^\ will be quite small at low values
of r, and the same holds for other proteins. Such a situation can occur during
emulsion formation. At the very short time scales involved, TT may even be smaller at
low r, because the molecules have not yet had time to expand in the interface and
thereby increase the value of ;r^\ Altogether, the modulus will be very small for
polymers at quite low F values.
We should add that the rigidity of a polymer coated interface may be higher if the
inter facial deformation is not truly dilational, but also has a shear component. This
has been observed for globular proteins and bituminous substances. Furthermore, the
equations and results given for \K^\ are only valid for quite small values of the strain.
During emulsification, very large strains and very high strain rates can occur, and the

' J. Benjamins, E.H. Lucassen Reynders, loc. cit.


^^ J A. de Feijter, J. Benjamins, J. Colloid Interface Set 40 (1982) 289.
EMULSIONS 8.15

modulus then is highly non Newtonian. Very little is known about the dilational
behaviour under such conditions.

8,Id Emulsion Properties


Several characteristics of emulsions have been mentioned or briefly discussed in
the previous sections. Some other properties are discussed below.

(i) Viscosity^\ In table IV.6.4 a collection was given of semi empirical formulas for
the (p dependence of the viscosity of dispersions of solid particles. Of these, we repeat
the Krieger-Dougherty equation [IV.6.9.10], which is also often used for emulsions:

r]=r]^{\-(p/(p^^)-^'J^^max [8.1.6]

where (p^gj^ is the volume fraction at which the viscosity will go to infinity. For
emulsions, where the particles are perfect spheres, the intrinsic viscosity [rj] would
equal 2.5. Possible complications are following:
(1) The effective volume fraction of the drops may be larger than that of the
disperse phase, because of the surfactant layer. If the latter has a hydrodynamic
thickness S, we have:

^^ff = ^(l + J / a ) 3 [8.1.7]

For a polymeric surfactant, J = 1 0 n m is a reasonable value, and for droplets of


a = 0.5 |Lim this results in (p^^^ - 1.06^ . In most emulsions the effect will be smaller.
(2) The fluidity of the disperse phase may contribute to (i.e. decrease) the emulsion
viscosity, according to the Taylor equation:
77^ +0.4/7p,
[77] = 2.5-^^ ^ [8.1.8]
^C + ^D

However, it is implicitly assumed that the viscous stress is continuous across the
drop's interface or, in other words, that Kg = 0 , see [8.1.5]. As discussed in sec. 8.1c,
under nearly all conditions the value of Kg will be large enough to prevent any lateral
movement of the interface, and hence any flow inside the drops. The drops will thus
behave as solid spheres.
(3) A drop can be deformed by the shear stress acting upon it in the rheometer,
which would increase the viscosity. As mentioned in sec. 8.1a, the deformation is
negligible under most conditions. See sec. 8.2.6, sub (i), for the factors determining
the deformation.
(4) Colloidal interactions between the drops will, in principle, affect the value of ;;.

Primarily based on a review on emulsion rheology by E. Dickinson, in B.P. Binks, Ed.,


Modern Aspects of Emulsion Science, chapter 5, p. 145. Royal See. Chem., Cambridge, 1998.
8.16 EMULSIONS

^D /^C

Figure 8.5. (a) Apparent shear viscosity 77^ versus shear stress, a, for emulsions (mineral oil
in water, d^2 = 0.55 fim , stabilized by Tween 20) of various volume fractions, (p, (indicated
near the curves, (b) Relative zero shear rate viscosity as a function of (p for; (1) SDS stabilized
o/w emulsions, dr^2 = 0.44 |j.m ; (2) the same emulsions as in (a); curve (3) is calculated
1)
according to (8.1.6] with (p^ = 0.71. Experimental results after

As will be discussed in sec. 8.1g, the effect is quite small in most emulsions, unless
the net attractive forces are large enough to cause permanent aggregation of drops.
Figure 8.5 shows some representative experimental results. Up to ^ ~ 0.4, the
Krieger Dougherty equation seems to be obeyed well, although a fairly small, correc-
tion for point 1 mentioned above would be needed. At higher <p values, deviations
from [8.1.6] are observed. The behaviour is no longer Newtonian and becomes
increasingly shear stress thinning, and the apparent viscosity (especially when extra-
polated to zero shear stress) increases much more strongly with (p than according to
[8.1.6]. Presumably, the observed deviations from the equation are primarily due to a
combination of points 1 and 4 mentioned above, in the form of an interaction range:
the distance from the 'bare' drop surface at which the drops sense each other. The
effect would be stronger for a higher ratio of the interaction range over the distance
between drops. The minimum distance between drops is about 0.05 times the radius
at (p = OA, and strongly decreases with increasing (p. The deviations will thus be
greater for a smaller drop size and a larger volume fraction. The nature of the surfact-
ant and the surface excess concentration can also influence the result.
The situation is different if the value of cp is larger than (p^^ for spheres. This has
been especially studied for foams, although most experimental results have been
obtained on emulsions-^^ Now the drops have deformed each other and the deform-
ation will increase upon application of a stress. This means an increase of the Laplace
pressure, which is the primary cause of the system now having an elastic modulus.

E. Dickinson, M.I. Goller, D.J. Wedlock, J. Colloid Interface Sci. 172 (1995) 192.
2)See e.g. the review by H.M. Princen, in R.D. Bee et al., Eds., Food Colloids, chapter 2, p. 14,
Royal Soc. Chem. (1989).
EMULSIONS 8.17

Semi empirical theory has been developed, in which the numerical constants have
been found by fitting to experimental results. For the elastic shear modulus we have

[8.1.9]
PL,a = 4 r ^ ^ / d 3 2

where the apparent average Laplace pressure, PL a' ^'^ calculated from the (assumed)
values for undeformed drops. It is seen that G vanishes at ^ < 0.71, but this value will
presumably vary somewhat with the shape of the droplet size distribution. The
modulus increases steeply with increasing (p .
Above (p- 0.72 the emulsion will have a yield stress, a , and for (7> G the emul-
sion will exhibit flow. Expressions for the yield stress and the apparent (shear rate
thinning) viscosity have also been given, but these are less certain. In principle, other
variables will contribute to the results. These include the surface dilational modulus of
the o-w interface and, for the viscosity, the value of rf^ . These aspects seem to have
been insufficiently studied.
Furthermore, the viscosity will be enhanced by aggregation of the drops, and by
addition of materials, especially polymers, that Increase the viscosity of the continuous
phase. In these cases, the emulsion will often show visco elastic behaviour.

(ii) Optical properties. The optical appearance of an emulsion is generally


dominated by the droplets' light scattering. The scattering is also of importance as a
tool in the estimation of droplet size distribution (sec. 8.1e), or interactions between
droplets (sec. 8.1g). Light scattering by small particles is discussed in 1.7, especially
sec. 7.8.
Small droplets scatter light according to the Raylelgh-Debye theory, which implies
that the shorter the wavelength, the stronger the scattering. For larger drops, say
> l | i m , we approach the domain of anomalous diffraction; the scattering then is
dominated by Interference between light passing through and light passing by a
particle . The scattering coefficient is given by,

4 4
P =2 sinp + —^(1-cosp)
p p^^ ' [8.1.10]
p = 27rd\An\/1

where p is called the phase shift parameter; An is the difference in refractive Index
between droplet and continuous phase, and A^ is the wavelength of the incident light
in vacuum. The theory is approximately valid for p > 2 and |An|<0.15 (accuracy
better than 10%).

^^ H.C. van de Hulst, Light Scattering by Small Particles, Wiley, (1957); there is also a Dover
edition, Dover (1981).
8.18 EMULSIONS

The total scattering Intensity is given by the turbidity,

_ 3;r^|An|g
[8.1.11]

which has dimension [ L"M- The absorbancy of a dispersion is given by 0.434 r^" ,
where t is the optical path length. Accurate values of Q can be calculated by the
rigorous Mie theory, which applies for spheres of arbitrary size and refractive index.
An example of r/(p versus p is given in fig. 8.6.

1.5h
T/(p
Figure 8.6. Turbidity over volume frac-
l.Oh tion (in |im~^) versus phase shift
parameter p for homodisperse spherical
particles, calculated for |Ari| = 0.1.
0.5h

The appearance of an emulsion is due to diffuse reflection, a complicated multiple


scattering phenomenon. Nevertheless, it correlates quite reasonably with the turbidity,
provided that t [i.e., the thickness of the emulsion layer), is not very small (say, over a
few mm).
For white light, i.e. average X^ = 0.55 |xm, we may conclude that for particles of
various diameters, the appearance of an emulsion will be:
d = 0.02 |im gra)ash, almost trainsparent
d = 0.2 |im bluish, transmitted light red
d =2 \im white
d = 20 i^m less white, maybe some colour

The smallest particles scatter very little light (Rayleigh scattering). Those of about
0.2 jLim scatter more, and the intensity depends strongly on the wavelength (Rayleigh-
Debye scattering), although the resulting colours will not generally be conspicuous
because of the polydispersity. In accordance with fig. 8.6, the maximum whiteness will
be observed for about 2 |im droplets. For still larger drops, the scattering is less and
the whiteness even more so, because a substantial part of the scattered light is
scattered at very small angles from the incident beam. Furthermore, total scattering is
proportional to particle concentration. (For very small drops and high (p the intensity
is diminished owing to dependent scattering.)
For droplets having a complex refractive index, fi = n-in\ absorption occurs
EMULSIONS 8.19

besides scattering, and the emulsion may show colour. However, for nearly all
emulsion droplets n ' < O . O l | n - l | for any 'visible' wavelength, and such a relatively
small absorption of light means that only for quite large particles can a colour possibly
be seen, since the total scattering is relatively small. If the continuous phase has a
colour, this is to some extent masked by scattering, especially for d ~ 2 |im and high
volume fraction.

[Hi) Dielectric properties. See sec. II.4.8 for a general introduction to dielectric
phenomena.
Although dielectric spectroscopy of emulsions has been studied, and provides some
interesting information, the interpretational problems are considerable and applica-
tion of the method is largely empirical. The central challenge is that of polarization
and permeability of the, generally charged, o-w interface.
Earlier studies emphasized static dielectric permeabilities of dispersions, often of
uncharged particles. Such studies led mostly to equations of the type e = fieico), e^, (p)
where co is the frequency of the electric field. Such equations emphasize the polariza-
tion of the two phases. Corrections for the interfacial contributions were sometimes
introduced by assuming the droplets to be surrounded by a shell of a different
dielectric constant. For reviews of these models see^\
More general approaches involve dynamic measurements in which the complex,
frequency dependent conductivity K[co) is measured and decomposed according to
[II.4.8.1],

k.[tm.,(o) = K^{co)-\-coe^[£'\Gvci,co]-e\^v[i,0))\ [8.1.121

where K^ is the conductivity of the continuous phase and 'em' refers to the entire
emulsion. For o/w emulsions, K^ is high; for w/o emulsions, it is low. The trend is
that with increasing co: K increases, e' decreases and ^" goes through a maximum;
some illustrations are in figs. II.4.37 39. The decrease of £\co] has been analysed in
some detail for dilute sols of charged spherical particles in electrolyte solutions. The
presence of electric double layers leads to gigantic values oi e' at low frequencies; with
increasing co the value of e' relaxes at a frequency that depends, inter alia, on the
particle radius. Models studied include latices and microemulsions. For true
emulsions, the situation is different, because; (i), they are rarely dilute; (ii), they are
generally polydisperse; and, (iii), depending on the type, polarization of the disperse
phase must also be considered. Condition (ii) implies a spectrum of relaxation times,
hence broadening of the £\(o] decay and the £'\co] maximum. Dielectric relaxation
measurement is not a suitable method for determining particle size distributions.

T. Hanai, in P. Sherman, Ed., Emulsion Science, chapter 5, p. 353. Academic Press, (1988);
and M. Clausse, in P. Becher, Ed., Encyclopedia oj Emulsion Technology, Vol. 1, chapter 9,
p. 481, Marcel Dekker (1983).
8.20 EMULSIONS

Condition (iii) is challenging: the permeability of ions, migration across and tangential
to the interface of ions and surfactants, slip phenomena, etc., are all reflected in €'{co]
and e'\co). Models to implement this would be needed.
Awaiting such general studies, most investigations have a semi empirical nature.
For instance, o/w emulsions stabilized by lecithins were studied^\ The £[co) spectrum
could be interpreted by four parameters, including a storage and a loss contribution of
the interfacial layer, which enabled the identification of lecithin preparations under
some conditions.
At high frequencies, only the continuous phase permittivities remain, which differ
between oil and water: relaxation in this range can help detect the oil and water
contents. This procedure is also in use for oil prospecting in rocks.

8.1e Determination of drop size distribution


Virtually all emulsions are polydisperse, so we need to know the size distribution^ .
The cumulative distribution of the droplet diameter, d, will be called F{d); it gives the
number of droplets of size smaller than d per unit volume, versus d; the dimension is
[ L~^ ] (where L stainds for length). The size frequency distribution is then defined as,
/ ( d ) = dF(d)/dd, of dimension [ L~^ ]. Often, the volume frequency distribution is
used, given by {7r/6)d^f[d), dimension [L~^].
A useful auxiliary parameter is the n th moment of the distribution, given by

Sn = j d V ( d ) d d [8.1.13]
0

which has dimension [L""^]. See also IV.app. 1. The quantity SQ gives the total
number of droplets per unit volume, 7182 the specific surface area A^ of the droplets,
etc. Any type of average droplet diameter can be expressed by;

d.,^{SJS,r^-^^ 18.1.141

which has dimension [L]. For the same distribution, an average type of higher order
(a + b) will have a larger value. The number mean diameter is given by d^Q , etc. Often,
the average type dgg (also called the volume-surface diameter d^^, or Sauter mean
diameter) is used, in part because of its relationship to the specific surface area; see
[8.1.1].
The spread in size is best expressed by a relative standard deviation.

^' R.M. Hill, E.S. Beckford, R.C. Rowe, C.B. Jones, and L.A. Dissodo, J. Colloid Interface ScL
138 (1990)521.
2)
An overview of distribution functions used for emulsions is by C. Orr in P. Becher, Ed.,
Encyclopedia of Emulsion Technology, Vol. 1, chapter 5, p. 369, Marcel Dekker (1983).
EMULSIONS 8.21

SS ^"'
[8.1.15]

For example, C2 is the relative standard deviation of the surface weighted droplet size
distribution, and so it corresponds to the average d^2 • ^^ ^^^ distribution is not too
wide, say, C2 < 0.5 , the distribution is usually close to log normal and can be
characterized with sufficient accuracy by the parameters d^2 ^^^ ^2 • ^^ ^2 ^^ l^rge,
the whole distribution should be given; it may have various kinds of shape, e.g. with a
long 'tail' or even bimodal.

Although the droplets in an emulsion are generally homogeneous spheres, reliable and
precise determination of the size distribution is far from easy. When applying two
different methods, it is not exceptional if the values obtained for a given type of average
diameter differ by a factor of two; the distribution shapes obtained may also differ
substantially. Some of the methods used are listed in table 8.1 ^^
The experimental methods vary widely in principle. Simplest is the determination
of the diameter of a large number of single droplets by microscopy, both light and
electron microscopy, applied in various modes of specimen preparation. Currently,
image analysis is often applied. Very useful for emulsions is confocal scanning laser
microscopy (CSLM), where the image is obtained directly in digital form; when using a
fluorescence microscopy mode, the limit of resolution is quite good. If the specimen is
liquid, it is generally advisable to suppress particle motion (sedimentation, Brownian
motion, convection currents) by the addition of a transparent gelling agent.
Some methods detect individual particles. The highly diluted emulsion flows
through a small sensing zone, where individual particles are sensed and sized. Best
known are the 'Coulter counter' and comparable instruments, where the particle
increases the electrical resistance in a narrow opening; the change in resistance is
proportional to particle volume. In another method, the amount of light scattered by
each particle is monitored; this needs a calibration curve to convert scattering
intensity to particle size^^ These methods tend to be relatively accurate. A problem
may be the reliability of the correction for coincidence, which is needed because two,
or even more, particles may simultaneously pass the sensing zone.
In most other methods a macroscopic property is determined, often as a function
of some external variable, from which the data are converted into an average size or a
size distribution. Most popular are light scattering methods, nearly always of strongly

A general book on methods is T. Allen, Particle Size Measurement, 5^^ ed., Vol. 1. Chapman
& Hall, (1997). Specifically for emulsions, C. Orr, in P. Becher, Ed., Encyclopedia of Emulsion
Technology, Vol. 3, chapter 3, p. 137, Marcel Dekker (1988).
^^ See e.g. E.G.M. Pelsers, M.A. Cohen Stuart, and G.J. Fleer, J. Colloid Interface Sci. 137
(1990)350.
8.22 EMULSIONS

Table 8.1 Summary of methods for the estimation of droplet size distributions
Method Determ- Droplet Dilution Problems/remarks
ines size range needed
(|^m)
Microscopy
Light microscopy fid) > 0.3 Generally Tedious, several pitfalls
CSLM fid) > 0.1 Generally Deconvolution; immob-
not ilizing may be needed
Electron microscopy fid) > 0.003 No Artefacts; 'tomato salad'
analysis

Particle sensor
Electrical resistance Fid) 0.5-300 Greatly Coincidence correction
Light scattering fid) 0.2-10 Greatly Tedious; can handle
droplet aggregates

Static light scattering


I Forward scattering Ii0) 0.4-200 Greatly Matrix inversion
Spectroturbidimetry r(A) 0.2 10 Yes Tedious, (p must be
known

Dynamic light scattering


Photon correlation /(At) < 1 Yes 'Average' only
spectroscopy

Nuclear Magnetic Resonance


Pulsed fif Id gradient Diffusion 1-10 No Rather tedious; no full
proton rfMR* time distribution

1 Sedimentation
Under gravity <^53 2-100 Generally Several
Centrifugal ^53 0.1-5 Generally Several

Note*: especially for w/o emulsions.


Symbols: / = intensity; 6 = scattering angle, r = turbidity; A — wavelength; At = time shift;
(p = volume fraction.

diluted emulsions. Often, forward scattering of laser light is measured over a range of
scattering angles. By use of Fraunhofer diffraction theory (for large particles) or of Mie
scattering theory (valid for spheres of arbitrary size and refractive index), it has been
tried to convert the angular spectrum into a size distribution. The available instru-
ments are very easy to use, and they immediately produce a size distribution, as well
as some distribution parameters. However, although the software for the methods has
gradually improved in quality, th« results can still be insufficiently reliable, especially
in the small particle range. Another method is spectroturbidimetry: the turbidity is
EMULSIONS 8.23

measured over a range of wavelengths, and the spectrum obtained is compared to


spectra calculated for various size distributions . The volume fraction and the refract-
ive indices (as a function of wavelength) must be known. The method is somewhat
tedious but gives reliable results, especially for droplet sizes of order 1 jim . Approx-
imate values of (^32 > 3 |im can be obtained from measurements of turbidity over cp at
one (relatively short) wavelength.
Besides these static scattering methods, dynamic (or quasi elastic) light scattering
can be useful^^ basic aspects are in sees. 1.7.6-8. Application to very dilute systems
will be considered first. Because the particles show Brownian motion, the wavelength
of the scattered light is slightly affected, whose shift is detected somehow. Various
modes exist, although photon correlation spectroscopy is now the method of choice.
Anyway, what is essentially measured is the diffusion coefficient of the particle, which
can, via the Stokes-Einstein relation, be converted into an equivalent sphere diameter.
In this way, the size of homodisperse small particles (d < 1 |im) can be estimated
accurately. For polydisperse emulsions, the problem is that the scattering intensity is
strongly dependent on size, so the diameter obtained is an average of high and un-
known order. Neither can the width of the distribution be established with accuracy.
More sophisticated methods, using a range of scattering angles, are in principle more
powerful'^^ Concentrated systems can be investigated by various forms of fibre optics
quasi elastic light scatterings^ one form being known as, 'diffusing wave spectroscopy'.
A problem is that at low (p the signal to noise ratio is too low, whereas at high (p the
diffusivity is affected by particle concentration and interactions. Hence, measurements
must be done at various cp and extrapolated to zero.
Pulse field gradient NMR methods can be applied to determine the size of droplets
by estimating the maximum distance over which the molecules of the material in a
droplet can diffuse. If proton NMR is used, the method is quite suitable for w/o
emulsions^ ^ It should be realized that most methods that demand strong dilution of
the emulsion are not, or are poorly, suitable for w/o emulsions.
Sedimentation analysis can yield, from the initial sedimentation rate, the value of
dgg, provided that the Stokes equation is obeyed; as will be discussed in sec. 8.3d,
several conditions (e.g., very small (p) must be fulfilled. By using gravity or centrifugal
sedimentation, a wide range of d^g values is accessible. Sedimentation can be com-
bined with some method for determining local droplet concentration, e.g., by

^^ P. Walstra, J. Colloid Interface Set 27 (1968) 493.


A clear introduction, including possibilities and limitations of various methods, is given by
D.S. Home in E. Dickinson, Ed., New Physlco Chemical Techniques for the Characterization
of Complex Food Systems, chapter 11, p. 240, Blackie (1995).
^^ See e.g. D.S. Home, D.G. Dalgleish, Em. Biophys. J. 11 (1985) 249.
Application to emulsions is discussed by P. van der Meeren, M. Stastny, J. Vanderdeelen, and
L. Baert, Colloids Surfaces A76 (1993) 125.
J.C. van den Enden, D. Waddington, H. van Aalst, C.G. van Kralingen, and K.J. Parker, J.
Colloid Interface Set 140(1990) 105.
8.24 EMULSIONS

ultrasound attenuation, to give immediate results. Figure 8.24 illustrates some


sedimentation profiles.
Some other methods are not (yet) very suitable for emulsions. These include, for
example, size exclusion or hydrodynamic chromatography (d < 1 )j,m); field flow frac-
tionation in a centrifugal field (d<l^im); electro acoustic methods; and ultrasound
velocity measurements.

Accuracy. Several kinds of uncertainty can arise in the determination of size


distributions. Systematic errors can occur readily for indirect methods. The signal
measured can depend on factors other than pairticle size. The relationship between the
magnitude of a signal and the particle size may not be known with sufficient accuracy;
often a linear relationship is assumed, but is not always true. Several methods under-
estimate the number of the smallest particles, or do not even notice them. This means
that the average size is overestimated, especially for averages involving SQ , such as
djQ and d^Q; an estimate of d^2 will then be closer to reality.
Even for direct methods, such problems may exist. Several microscopic methods,
in fact, see cross sections or thin slices of the material. Then a number of circles is
observed, and the problem is to convert their diameter distribution to that of the
original spheres; this is known as the 'tomato sadad problem'. Good solutions exist for
spheres, which are to be found in texts on stereology^^ The droplets are often allowed
to sediment (or cream) before they are viewed, and the small ones may then escape
notice. Furthermore, it may be difficult to distinguish between separate particles that
are close to each other and previously formed aggregates, especially when using an
image analysis programme. In indirect methods, the presence of aggregates may
completely upset the results, and care should be taken that the droplets are fully
dea^regated.
Conversion of the raw data to a size distribution especially poses problems for
indirect methods. For instaince, in scattering methods a range of data (a spectrum) has
to be determined, be it a range of wavelengths or a range of scattering angles, to allow
derivation of amything more than an average size. If the particle size distribution and
the refractive indices are known, it is relatively easy to csdculate a spectrum. However,
the inverse problem, calculating the distribution from a spectrum, is far more difficult,
especially because the amount of information and its accuracy are limited. The
algorithms involved always involve rounding off, or even shortcuts, and may even be
based on a given type of size distribution. All of this may lead to considerable error.
Finally, the reproducibility should be taken into account. Sampling should be
representative. Sizing may give rise to random errors. The greatest uncertainty is
generally due to random errors in the counting or, more precisely, in establishing the
number of particles in a size class. This is largely owing to the Poisson statistics of

^^ See e.g. E.R. Weibel, Stereological Methods, Vols. 1 and 2, Academic Press, (1980).
EMULSIONS 8.25

counting. The standard deviation of the number of particles in a certain volume is, for
a completely random distribution, equal to the square root of the average. This means
that the relative standard deviation of the number of particles in a size class, i, is
equal to, or larger than, 1/JiV" , where AT. is the number actually counted [i.e., before
any multiplication with a dilution factor, etc.). Counting just one particle thus leads to
an uncertainty (relative standard deviation) of over 100%. This becomes especially
manifest for very large particles, which means that the large particle end of a volume
distribution is often subject to considerable error. Especially if the size distribution is
wide (high value of C2 ), tens of thousands of particles may have to be counted to
obtain reliable results.

Another variable of great importance is the total volume fraction [^=(71/6)83] of


the droplets. For some methods of estimating size distributions, (p can be derived
from the results, but it is better to use this as check, after having obtained the value of
<p In another way. Deriving it directly from the quantities of materials used in making
the emulsion is often inaccurate, especially when making small volumes. Chemical
analysis of the components making up the disperse phase, or of an added marker
substance that only dissolves in the disperse phase, is to be preferred.

8, If Determination of interfacial properties


This section concerns interfacial tension, interfacial rheology, surface excess
concentration and surface layer composition.
Nearly always, interfacial properties are determined at a macroscopic interface. The
continuous phase as used in making the emulsion, i.e., with surfactant(s) added, and
the material making the disperse phase, are brought into contact and measurements
are made on the interface. These may concern static or dynamic interfacial tensions,
and surface rheology (see Vol. Ill, sees. 3.7e for Langmuir monolayers, and 4.5 for
Gibbs monolayers). However, the macroscopic monolayer can be representative of the
monolayer around the droplets in only one case, when there is one (pure) non
polymeric surfactant, which is insoluble in the disperse phase, and its concentration
in the emulsion is in excess of the concentration that provides a plateau value of / . In
all other cases, the composition of the macroscopic monolayer will not equal that
around a droplet in the emulsion. This is because the surface area to volume ratio is
higher in the emulsion than in the macroscopic case by some orders of magnitude,
and because some of the surfactant species will be depleted more strongly than others
during emulsification. Moreover, it may take a long time to obtain equilibrium. See
also sees. 8.1b and c.
What can be done, is to make the emulsion and, after allowing some time for
equilibration, separate the continuous phase by centrifugation or microfiltration. This
liquid and the material of the disperse phase are brought into contact, providing an
interface. After keeping the system for a considerable time (some hours often suffice)
8.26 EMULSIONS

the interfacial composition may be virtually equal to that in the emulsion, provided,
again, that the surfactants do not dissolve in the disperse phase. Moreover, the
separated liquid must be free of droplets. For some polymeric surfactants, notably
proteins, the monolayer composition in the emulsion may again be different, because
an equilibrium interfacial composition may not have been reached; see sec. 8.2c.
In some cases, the interfacial tension can be measured at the surface of a droplet,
provided this is larger than, say, 4 jim. Under a microscope, a drop in an emulsion is
partly sucked into a very narrow cylindrical capillary, and the pressure difference to
achieve this situation is established. This must be equal to the difference in Laplace
pressure, given by 2y{l/R^ - I / K 2 ) . Here, Rj is the radius of the capillary and ^2 that
of the protruding part of the drop, and these radii can be measured. In this way, the
interfacial tension can be obtained^\ It is a very tedious method, but can be used if
needed to check results obtained at a macroscopic interface.
Altogether, reliable data on interfacial tension and rheology of monolayers on emul-
sion drops are hardly available.

Estimation of the surface excess F is, in principle, not difficult, although it is not so
easy to obtain high accuracy'^^ One determines in the emulsion the values of (p, d^2
(which yield A^ =6(p/d^2^ ^^^ ^he concentration of surfactant, c, in moles or mass
units per unit volume. We now have

c = c^{l-(p) + 6r(p/d^2 [8.1.16]

where c^ is the concentration in the continuous phase. After separating most droplets
from the emulsion by centrifuging or microfiltration, the same values are determined
in the 'skimmed' emulsion. We now have two linear equations with the two unknowns,
r and c^, which can therefore be calculated. The assumptions underlying this
method are that no desorption of surfactant occurs during separation, that the
surfactant does not dissolve in the disperse phase, and that F does not depend on the
droplet size.
Often a simpler method is tried: one determines c^ in a skimmed emulsion in
which (p is virtually zero, which yields the amount of surfactant adsorbed. However,
this is not really possible if the emulsion contains a significant amount of very small
drops, say < 0.4 |Lim .
The composition of a mixed adsorption layer can be obtained by determining F for
each surfactant present. Some polymers, especially proteins, effectively give a
Langmuir monolayer. In such a case, one can 'wash away' the dissolved surfactant by
repeated dilution and skimming of the emulsion. One may then desorb the adsorbed

^' L.W. Phipps, D.M. Temple, J. Dairy Res. 49 (1982) 61.


^^ H. Oortwijn, P. Walstra, Neth. Milk Dairy J. 33 (1979) 134.
EMULSIONS 8.27

polymer by adding a suitable amphiphile, £ind remove the disperse phase material
(generally oil). The poljrmer solution can then be amalyzed by a chemical method^^ If
the emulsion drops are small, and cp is not high, addition of an excess of SDS not
only dissolves the protein: the oil is adso 'solubilized' in SDS micelles. The resulting
liquid can be analyzed directly by SDS gel electrophoresis ^^

S.lg Determination of colloidal interaction forces


Knowledge of the colloidal interaction Gibbs energy (G), or the corresponding force
(F), between emulsion droplets as a function of inter droplet distance, (h), will be
very useful, since these are essential parameters in determining whether aggregation
and/or coalescence of drops will occur. Pair interactions between hard surfaces are
discussed extensively in ch. IV.3, which also gives methods of determination. We shall
now discuss the extent to which these methods also work for emulsions.
Nearly adl measurements are done on macroscopic surfaces, e.g., determining the
disjoining pressure in a liquid film in various states of drainage, or measuring directly
forces (in the surface force apparatus), as a function of h, between very smooth solid
surfaces covered by surfactant monolayers. The results are generally not represent-
ative for emulsions. In the first place, the monolayers involved may differ in composi-
tion from those on the emulsion drops, because, (i), the adsorbent is generally
different (a-w or s-w, rather than o-w) and, (ii), the monolayer will generally not have
the same composition, for reasons given in the previous section. Nevertheless, some
interesting results have been obtained, which do throw light upon the colloidal inter-
actions between emulsion droplets.
Secondly, the time scales involved can be quite different. Measurements on macro-
scopic surfaces take at least severad minutes. The time during which two emulsion
drops are close to each other is far shorter. Consider droplets of diameter 1 |Lim in
water. The time, t, needed for a particle to diffuse in a given direction over a distance
zlj = 2yfDi , where D is the diffusion coefficient. From the Stokes Einstein relation-
ship, we find for an individual droplet, D = 4 x l 0 " ^ ^ m ^ s " ^ . The pair diffusion
coefficient will be twice as large, but at close approach the diffusion will be greatly
slowed down. When, for example, we consider diffusion from 25 to 5 nm interparticle
distance, it is by a factor of about 100 ^\ Hence, we will have D == 10"^"^, and the time
needed for diffusion over 20 nm would be about 10 ms. This is shorter by some
orders of magnitude than the time needed to obtain an equilibrium composition for a
mixture of surfactamts, and an equilibrium conformation of adsorbed polymers.
Consequently, it would be desirable to make dynamic measurements directly on
emulsion droplets. One method, called 'colloidal particle scattering' has been

^^ J.A. Hunt, D.G. Dalgleish, Food Hydrocoll. 8 (1994) 175.


^^ See e.g. L.A. Spielman, in K.J. Ives, Ed., The Scientific Basis of Flocculation, Sijthoff &
Noordhoff(1978)p. 65.
8.28 EMULSIONS

Figure 8.7. Principle of the colloidal particle scattering apparatus, (a) Shown are a fixed
particle (F) and a mobile one (M) in its original position, as well as the trajectory of the latter in
the z-y plane, (b) A cross section through F in the x-z plane (seen in the y direction), v =
liquid velocity, a = particle radius, R and 6 are the cylindrical coordinates of the centre of M.

developed by van de Ven and co workers^^ The principle is illustrated in fig. 8.7. On a
plate is fixed a particle and a dilute dispersion of the same particles flows past it, in
simple-shear. The trajectories of several particles that come close to the fixed particle
are monitored via a light microscope. These particles will be deflected and show a
trajectory roughly as depicted, which is calculated for the case where there is only
hard core repulsion between the particles. The trajectories before and after the
encounter are considered. The coordinates of the moving particle will have changed in
the X and the z directions. When using cylindrical coordinates, as depicted in fig.
8.7b, it is generally observed that the angle 6 changes very little; hence, the deflection
can be expressed in the change in the radial coordinate, AR. Let AR^ be the (cal-
culated) deflection if only hard core repulsion and no attraction occurs. When colloidal
repulsion acts between the particles, e.g., of electrostatic or steric nature, the result
should be that AR > AR^ . For net attraction, the opposite is to be expected. In fact, the
static particle may capture the moving one if the attraction is strong, but the latter
generally escapes after rolling for some distance over the former one, owing to the
presence of the supporting plate. Nevertheless, the proportion of encounters that lead
to capture can be established.
One can try to estimate G[h) by comparing observed trajectories, or merely AR
values, with those calculated for various assumed relationships between G and h .
This is no simple task. First, the calculation of the trajectories is intricate, time con-
suming, and not completely exact. Second, there are some important disturbances,
since the particles are also deflected by sedimentation (in a single direction) and by
Brownian motion (in a random way); the latter means that the translational Peclet
number for the particles (shear force over the diffusional force, i.e. cfiVv/SD, where

^'T.G.M. van de Ven, P. Warszynski, X. Wu, and T. Dabros, Langmuir 10 (1994) 3046. We
shall largely follow papers from the group of Dickinson, especially M. Whittle, B.S. Murray, E.
Dickinson, and V.J. Pinfield, J. Colloid Interface Set 223 (2000) 273; and M. Whittle, B.S.
Murray, and E. Dickinson J. Colloid Interface Sci. 225 (2000) 367.
EMULSIONS 8.29

Vf is the shear rate, and D is the particle diffusion coefficient) should be high, say
> 1 0 0 . Third, the accuracy needed may be higher than the apparatus can provide;
moreover, the particles should be perfect spheres and not too small (say, > 4 fim ) to
obtain reliable results. This means that the method is not suitable for routine work.
On the other hand, some interesting results have been obtained, although the experi-
ments often involved density matching to minimize sedimentation, and increasing the
viscosity to suppress Brownian motion; the inherent compositional changes may
therefore have affected the surfactant adsorption. Moreover, emulsion drops are not as
easy to handle as polystyrene latex particles, and the latter are often used as a model.
Although these may be good as an adsorbent for surfactants such as proteins, where
the monolayers mimic those on oil droplets, the van der Waals attraction between
these particles is inevitably different from that between oil drops (see e.g. IV.app. 3).
For emulsions stabilized by polymeric surfactant(s), it may be interesting to
estimate the hydrodynamic thickness of the adsorbed monolayer, since this parameter
correlates well with the range over which steric repulsion acts. If the average droplet
size is small (say, d < 0.6 |Lim ), this can be done by determining the droplet size by
photon correlation spectroscopy before and after removal of the pol5mier from the
surface by adding a suitable small molecule surfactant ^\

Some macroscopic 'solution' properties can also yield information on colloidal inter-
action forces, provided that no irreversible a^regation of particles occurs. One group
of methods employs light scattering (discussed in sec. IV.2.3). Static scattering in the
Rayleigh-Debye domain at a range of volume fractions results in a structure factor that
depends on colloidal interactions. It is limited, however, to quite small particles (say,
d < 0.6), which makes it unsuitable for most emulsions. For small particles, inter-
action forces can also be derived from dynamic light scattering, but the method is
rather unreliable for nearly all emulsions, owing to their moderate to strong poly-
dispersity.
Another method involves the determination of specific viscosity as a function of
volume fraction. For hard particles, the relationship is often given as;

%^=^^—^ = k^(p + k2(p'^+k^(p^+... [8.1.17]


^c

where 77^ is the viscosity of the continuous phase, see sec. IV.6.9a. For hard, non-
interacting spheres, /Cj = 2.5. For ^ < 0 . 1 5 , the third term is generally not needed.
The second term is due to volume exclusion of, and pair interactions between, the
particles. For hard non-interacting spheres, k2 =6 . The constants are different when
colloidal interaction forces act between the particles. A repulsive force primarily leads
to a (small) increase in k j , depending on the strength and range of the interaction. In

^' D.G. Dalgleish, Coll SurJ. Bl (1993) 1.


8.30 EMULSIONS

the case of attractive forces, /C2 is increased^\ In a common treatment, where the
attractive free energy is depicted as a narrow square well, ^2 would equal
(5.9 + 1 . 9 / r g ) , where r^ is the Baxter parameter, related to the second virial coef-
ficient for, say, osmotic pressure by B2 = 4 - 1 / 1 3 . For B 2 = 0 , r = 0 . 2 5 , and for
stronger attraction its value is smaller. The equation predicts results that agree well
with observations, as long as the spheres are not very polydisperse.
The treatment is based on the assumption that the translational Peclet number,

37ld^77pVu
Pe = -^— [8.1.18]
8/cT

is much smaller than unity, which needs quite small particles and low shear rates.
Ideally, T/Q should be determined, i.e., as an extrapolated value at zero shear rate. For
most emulsions, however, Pe » 1 . This implies that the constants in [8.1.17] will be
different, but the theory has neither been well developed, nor checked. IVIorcover, any
effect of colloidal interaction will be relatively weak, since hydrodynamic effects are
prevalent. This means, again, that interaction forces cannot be determined readily in
most emulsions.

Attractive forces leading to irreversible aggregation naturally give rise to aggregates


(coagulates or flocculates). This will greatly increase viscosity, owing to the Increase in
effective volume fraction, and also increase light scattering. This yields possibilities for
determining aggregation. However, one may also use a simple light microscope to
detect aggregates.

8.2 Emulsion Formation

Emulsion formation involves several aspects and can be achieved in several ways. The
literature on this subject is vast, and we will primarily follow the contents of three
reviews by the present author'^^ supplemented with newer literature.

8.2a Introduction
To make an emulsion, one needs oil, water, surfactant and energy. Generally, mechan-
ical energy is employed, but chemical or electrical energy can also play a part. Energy
is needed because the interfacial free energy of the emulsion is higher than that of the
original non emulsified mixture, by an amount A^y. The counteracting increase in
mixing entropy upon emulsification is negligible. If we assume that we make an
emulsion with ^ = 0.1 and d^2 =0-3 |im , and that the final value of ^ = 10 mN m"^,
the increase in Gibbs energy amounts to 20 k J m~^. However, the amount of

^' See e.g. A.T.J.M. Wouterse and CO. do Kruif, J. Chem. Phys. 94 (1991) 5739.
^^ P. Walstra, (1983, 1993), and P. Walstra, P.EA. Smulders (1998); see General References.
EMULSIONS 8.31

mechanical energy expended is likely to be about 20 MJ m~^ . This huge excess is to a


small extent, due to the droplet surface area and the y value being temporarily larger
than their final values; moreover, this occurs a number of times (say, 100) during the
process. However, by far most of the mechanical energy is dissipated into heat, since it
acts on all of the liquid during all of the time of emulsification. The process is thus
very energy inefficient.
Much research is thus aimed at increasing the efficiency. Moreover, it is often tried
to obtain narrow drop size distributions. The larger the droplets, the lower generally is
the emulsion stability , but the amount of energy needed is larger for smaller droplets.
The optimum is clearly a homodisperse emulsion, and methods have been developed
to achieve this, but have not yet met with much success for large-scale manufacture.

(i) Emulsification methods. A wide range of apparatus and process-conditions can


be applied to make emulsions. The underlying principles can be classified as follows.
1. Nucleation and growth^\ In some cases, the material intended to make up the
disperse phase is, to some extent, soluble in the continuous phase, e.g., water in
benzene. It may then be possible to realize supersaturation, e.g., by altering the
temperature or by partial evaporation of the continuous phase. This will then generally
cause nucleation, followed by growth, of the disperse phase. Fairly homodisperse
drops can often be obtained, and if needed, nucleation may be regulated by adding a
limited amount of tiny catalytic impurities. However, coalescence of the newly formed
drops becomes appreciable when d and (p increase, leading to polydispersity.
Coalescence can then be stopped by adding a suitable surfactant. Careful optimization
of the process conditions may result in a reasonably homodisperse emulsion, although
it will eventually be subject to strong Ostwald ripening (sec. 8.3b). The author is
unaware of large-scale application of the process.
2. Injection. When a liquid is forced through a narrow capillary, a cylindrical
thread emerges. The thread is subject to Raylelgh instability^^ and if the Reynolds
number Re « 1 and the inter facial tension is constant, the diameter of the drops will
be 2 to 3 times the internal capillary diameter, depending on the viscosity ratio of the
disperse over the continuous phase.
Based on this principle, membrane emulsification has been developed^^ Drop
formation is illustrated in fig. 8.8. The 'membrane' is a sheet of porous material. The
future disperse phase (say, oil) is pressed through the sheet; also here Re « 1 in the

For a review see B. Vincent, Z. Kiraly, and T.M. Obey, in B.P. Binks, Ed., Modern Aspects of
Emulsion Science, chapter 3, p. 100, Royal Soc. Chem. (1998).
^^ See Ill.fig. 5.47.
There is considerable literature on the subject, for the most part of a practical nature. A good
overview of all the variables affecting the results is by V. Schroder, Herstellen von Ol-in-Wasser-
Emulsionen mit mikroporosen Membranen, Shaker Verlag, (1999). Some fundamental aspects
are discussed by N.C. Christov, D.N. Ganchev, N.D. Vassileva, N.D. Denkov, K.D. Danov, and
P.A. Kralchevsky, Colloids Surf. A209 (2002) 83.
8.32 EMULSIONS

pores. Along the membrane is made to circulate a surfactant solution, the future
continuous phase:: the two phases are thus in cross flow. Oil 'drops' emerging from
the pores are dislodged from the membrane by viscous forces exerted by the water
phase. To allow dislodging, the membrane must be preferentially wetted by the
continuous phase. Ideally, all pores should have the same diameter, and be separated
from each other by a distance of at least three diameters, to minimize immediate
coalescence of emerging drops. If these conditions are fulfilled, and a suitable
surfactant is present in sufficient concentration, and the flow in the pores is relatively
slow, a roughly homodisperse emulsion is formed, with d ~ Sd^ .

Pi

continuous P2
phase -^ emulsion
t ^^'":;;•_•• ^ ^ ^ L ^ ^
disperse
phase

membrane

Figure 8.8. Membrane emulsification. Cross section through a cylindrical emulsification


module, p^ - P2 is the trans membrane pressure. Also, see text.

The trans-membrane pressure should be at least larger than the Laplace pressure
in the pore, i.e.,4y/d^, which ranges for the most part between 2 and 100 kPa. To
achieve a superficial oil flow rate (volume flow rate per unit ofmembrane surface area)
that is large enough for practical purposes, e.g., 10~^m s~^, a much higher pressure
has to be applied, e.g., by a factor of five. The result is that the drops are larger and
far less homodisperse. Factors leading to a smaller average droplet size are: a smaller
pore diameter; a lower trans membrane pressure; a higher viscous stress at the
membrane surface (given by T/VUQ where VVQ is the velocity gradient at the surface); a
higher concentration, and a more suitable type of surfactant. It appears that large
drops often result from immediate coalescence, especially when partial wetting of the
membrane by the drops occurs, which depends, in turn, on the type and concen-
tration of surfactant. Moreover, inertial forces may cause formation of larger drops at
a higher flow rate through the pores, also if no coalescence occurs. Finally, a longer
time lag before the drop is dislodged (for example, owing to a low flow velocity of the
EMULSIONS 8.33

continuous phase) causes an increase in drop size, possibly owing to enhanced


coalescence. The time scales for drop formation are relatively long, for the most part
between 0.1 and 10 s. The results also depend significantly on the structure of the
membrane. A full explanation of the factors affecting the process has not yet been
given.
Membrane emulsification offers a good method for making homodisperse emul-
sions (droplet diameter > 1 ^im) on a laboratory scale. Larger scale production may
become possible and profitable for small cp (say, < 0.1) and dgg > 2 |j,m ; the width of
the droplet size distribution is substantial. The advantage is that the amount of
mechanical energy that is dissipated into heat can be substantially smaller than in
most other methods. Current research may lead to further improvement.
3. Aerosol to liquid. One may atomize a liquid in air by forcing it at very high
velocity through a nozzle. A spray of fine droplets can be obtained and be directed to
hit another liquid, to form sin emulsion. This is somewhat akin to injection, but the jet
leaving the opening has a high Re, and drop formation is due to inertial forces. The
droplet size distribution tends to be quite wide, and the method has no real advantage
over others.
A spray of fine droplets can also be obtained by applying a high electric potential
(up to +10 kV) to a liquid of high dielectric constant (water, ethanol, etc.). A receptacle
then should be placed not far away from the spraying point, and have a negative
potential. Small quantities of reasonably fine and homodisperse w o emulsions can be
obtained in this way^\
4. Ultrasonic emulsification^K An ultrasound generator causes great pressure fluc-
tuations £ind can thereby produce cavitation in a liquid that has a substantial vapour
pressure. For emulsification, ultrasound of frequency 20 kHz is generally used. At
high sound intensity, many very small cavities are formed. Such a cavity soon collapse
again and it does so over a very short time (e.g., within 10 ns), thereby inducing a
shock wave in the liquid. The wave concentrates, and loses energy, at sites where the
compressibility changes, hence at a phase boundary. The interface can be disrupted,
so drops are formed and disrupted again into smaller ones. The mechanisms involved
are compsirable to those in the TI regime discussed below. Small drops can result, but
the size distribution tends to be quite wide, and is often bimodal. Upon prolonged
exposure to ultrasound, a unimodal distribution of very small drops (down to
d^2 = 0 . 1 5 |im ) can result, although at the cost of much energy.
Cavitation can also occur in other machines, e.g. in some high pressure homog-
enizers. Intense cavitation may also degrade high polymers and even induce chemical
reactions.
5. Agitation. When a liquid is intensely agitated, hydrodynamic forces deform and

^^ See e.g., E.S.R. Gopal, in P. Sherman, Ed., Emulsion Science, Academic Press (1968) p. 1.
^^ See e.g., P. Walstra, (1983) loc. cit.
8.34 EMULSIONS

d i s r u p t fluid particles in the liquid. This is by far the m o s t applied m e t h o d for


e m u l s i o n formation. T h e a p p a r a t u s u s e d varies from s i m p l e s t i r r e r s or static m i x e r s
to m o r e s o p h i s t i c a t e d m a c h i n e s t h a t can p r o d u c e very intense agitation; see fig. 8.9. In
(a) a r o t o r s t a t o r type m i x e r is depicted t h a t is c o m m o n l y u s e d for low viscosity
l i q u i d s (it is often referred to by a t r a d e n a m e , e.g., Ultra T u r r a x ) . In (b), a colloid mill
is s h o w n , a r o t o r s t a t o r device for highly viscous liquids. A high p r e s s u r e homog-
enizer, (c), is u s e d for low viscosity liquids; the liquid is p u t u n d e r high p r e s s u r e a n d
forced t h r o u g h a very n a r r o w slit, w h e r e potential energy is first converted into kinetic
energy w h i c h is s u b s e q u e n t l y d i s s i p a t e d into heat. It r e s u l t s in very intense agitation
for a very s h o r t time, 0 ( 0 . 1 m s ) .
In t h e r e s t of sec. 8.2 we will primarily c o n s i d e r e m u l s i o n formation during
agitation. It m a y b e a d d e d t h a t surface forces arising from interfacial t e n s i o n g r a d i e n t s
can, in s o m e c a s e s , c o n t r i b u t e to d r o p l e t formation; see, e.g., sec. 8.2c, s u b (v).

(a] (b] (c)

stator^ I ^rotor
stator

rotor

Figure 8.9. Cross section of the active part of some emulsion making machines, (a) Rotor stator
type stirrer, (b) Colloid mill, (c) Valve of a high pressure homogenizer; the zig zag line denotes a
heavy spring. The slit width in (b) and (c) is greatly exaggerated. The arrows indicate the flow
direction. (After P. Walstra, Physical Chemistry of Foods, Marcel Dekker (2002).)

(ii) Phenomena occurring. Droplet formation owing to h y d r o d y n a m i c forces c a n


o c c u r by v a r i o u s m e c h a n i s m s ^ ^ We will not d i s c u s s t h e p h e n o m e n a involved, since
they a r e never critical: it is very easy to m a k e a c o a r s e e m u l s i o n . However, n o t all
m a c h i n e s c a n m a k e e m u l s i o n s : for example, a high p r e s s u r e homogenizer c a n only
h a n d l e a p r e m a d e (coarse) e m u l s i o n : w h e n a n o n emulsified m i x t u r e of oil a n d water
p a s s e s t h r o u g h the valve, the whole slit would b e filled either with oil or with water.
T h e critical step, then, is the b r e a k - u p of d r o p s into s m a l l e r o n e s . This n e e d s d r o p
deformation, which is c o u n t e r a c t e d by the Laplace p r e s s u r e ,

1 1
PL = y —+ — [8.2.1]
V^l Rn

See E.S.R. Gopal, loc. cit., and P. Walstra, (1983) loc. cit.
EMULSIONS S.35

where R^ and R2 are the principal radii of curvature of an interface. Deformation will
lead to an increase in p ^ , as is discussed, e.g. in ch. III.l. The resistance against
deformation is therefore greater if the drop is smaller and y is larger.
Drop deformation must be followed by its break-up into smaller ones, which leads
to an increase in surface area. Hence, surfactant must be transported to the aew
surface. Moreover, newly formed droplets will frequently encounter each other and
then may coalesce; whether this occurs depends, among other factors, on the
surfactant load. The processes are illustrated in fig. 8.10. They all have their own time
scales, as discussed below.

(a) .^o
(b)

(c)

(d)
DO-00 - 0 0
Figure 8.10. Various processes occurring during emulsion formation. Drops are depicted by
thin lines and surfactant by heavy lines and by dots. Schematic, and not to scale. The dichotomy
between fully covered and bare surfaces is also an oversimplification. (After P. Walstra and
P.E.A. Smulders, (1998) loc. cit.)

(Hi) Regimes. In agitated systems, the forces acting on the drops are of a
hydrodynamic nature. It is useful to distinguish between some emulsification regimes,
based on two important criteria:
Flow type (see sec. 1.6.4), being laminar (L) or turbulent (T), depending on the
Reynolds number. We recall that Re = Lvp/ rj, where v is the (average) liquid velocity.
The characteristic length, L , depends on the geometry: it equals four times the area
over the peripheral length of a cross section of the vessel perpendicular to the flow
direction. The critical Re for turbulence to occur depends on geometry, but is often
roughly 2300.
Type of force. Forces can be frictional, also called viscous (V), or inertial (I). Vis-
cous forces act tangentially to the droplet surface and are given by rjVv, where Vu is
8.36 EMULSIONS

Table 8 . 2 . Various regimes for e m u l s i o n formation. See text.


Regime LV TV TI
Flow type laminar turbulent turbulent
Forces viscous viscous inertial
Reynolds number:
for flow < 2000 > 2500 > 2500
around drop < 1 < 1 >1^^

„l/2„l/2 ^2/3^2/3^1/3
External stress acting on a drop TJcVu

r j,3/5
Resulting d ^
E^^lf^^ ^2/5^1/5

Characteristic times
HD ^D
^l/2„l/2 ^2/3^2/3pl/3
Mef ~
£ 'Ic

dm^Vv
6dm^e'^''
"^ads ~
^/^p..-

n d2/3p'/3
^enc ~ 8(p\/v 8(^^1/2

^^ Re^^ is > I if d> rj^/yp. Applies only if % » ^c-


Symbols: Re = Reynolds number; We = Weber number; d == droplet diameter; £ = power
density (energy dissipation rate); 7 = inter facial tension; 77= viscosity; Vu = velocity gradient;
T— characteristic time scale; r= surface excess; m = [surfactant]; ^ = volume fraction of
disperse phase; We and e are defined below.
Subscripts: D = disperse phase; C = continuous phase; def = deformation of a drop; ads =
adsorption of surfactant; enc = encounter with another drop; cr = critical value; pi = plateau
value.

t h e local velocity gradient. Inertial forces c a n b e d u e to velocity fluctuations t h a t lead


to fluctuations in p r e s s u r e (p), according to Bernoulli's law:

P + ^ P ^ ^ = constant [8.2.2]

Inertial forces act perpendicularly to the d r o p surface. Whether viscous or inertial


forces a r e p r e d o m i n a n t d e p e n d s on the d r o p Reynolds n u m b e r , Re^^., w h e r e L = d
a n d V is the velocity of the d r o p relative to the liquid. The critical n u m b e r for inertial
forces to occur e q u a l s a b o u t unity.
T a b l e 8.2 gives p a r t i c u l a r s of the relations in the various regimes. Besides the
criteria for the regime, the resulting s t r e s s , the a p p r o x i m a t e droplet size obtained, a n d
c h a r a c t e r i s t i c t i m e s for the various events occurring d u r i n g emulsification, are given.
In practice, i n t e r m e d i a t e situations occur; hence, the relations given are a p p r o x i m a t e .
EMULSIONS 8.37

Moreover, most of the equations are scaling laws, although the proportionality factor
is generally of order unity. Nevertheless, the equations are quite useful, if only because
they make clear what variables are essential. The underlying theory will be discussed
further below.
In all of the theories underlying the relations in table 8.2 it is assumed that the flow
is unbounded, meaning in practice that the drops formed are much smaller than the
width of the slit in which break-up occurs. However, in very small machines, especially
some laboratory homogenizers, this condition is not fulfilled, since d is comparable to
the slit width. This gives rise to a fourth regime, 'bounded flow' (LB) the flow is always
laminar (and the term agitation may be inappropriate). It will be briefly discussed
below.

8.2b Hydrodynamics
In this section, the interfacial tension is assumed to be constant. This means either
that no surfactant is present, which implies that a single drop is considered or that
the surfactant is in excess and conditions are such that r^^^ « r^^^ ; see table 8.2. In
the first two subsections, unbounded flow is assumed. A brief introduction into flow
types is in sec. 1.6.4; see especially fig. 6.7.
(i) Laminar Jlows^\ This concerns the Regime LV. Viscous forces are responsible
for deformation and break-up of drops. Examples are given in fig. 8.11. The external
stress, <j^^^, exerted by the flow equals r/Vv , where Vu is the velocity gradient (strain
rate) at the drop surface. For simple-shear flow it equals the shear rate: for true
elongational flows, the elongation (extension) rate. Here, ?] is the viscosity of the
surrounding liquid, but its value depends on the flow type. For example, in
elongational flow, 77^^ = Tr 77, where the proportionality constant is called the Trouton
number. Tr = 2 for uniaxial two dimensional flow (depicted in fig. 8.1 lb) and Tr = 3
for axisymmetric uniaxial flow; for biaxial flows, the values are 4 and 6, respectively.
The quantity rf is the 'common' shear viscosity, and this value is usually determined
and inserted in equations. However, the stresses acting on the drops are larger than in
the case of simple-shear flow. The numerical values given are only valid for Newtonian
liquids; for visco elastic liquids Tr will be (much) larger.
Calculation of drop deformation, under the assumption that the drop shape
remains very close to equilibrium during the deformation process, has been applied
successfully for various flow types. Use is made of a Weber number (by some authors
called the capillary number for laminar flows), the external stress divided by the
Laplace pressure, which for viscous forces is defined as

We = Wv7]^d/2Y [8.2.3]

^^ More information is in P. Walstra (1983), and P. Walstra and P.E.A. Smulders (1998): see the
General References. A review on the theory of, and experiments on, drop deformation and
break-up is given by H.A. Stone, Ann. Rev. Mech. 26 (1994) 65.
8.38 EMULSIONS

For simple-hear flow the deformation of the drop is given by, D = {L-B)/{L +B], as
iflustrated in fig. 8.11a. For W e « 1, D = We; incidentally, this imphes that the
elastic shear modulus of a drop equals 2 7 / d or half the Laplace pressure. It is seen
that the drop turns into a prolate ellipsoid, with the long axis at 45° to the flow
direction. The liquid inside the drop rotates, in accordance with simple-shear being a
rotational flow. The drop is broken up if We exceeds a critical value. The magnitude
of We^^ is given in fig. 8.12 (curve marked a = 0 ) as a function of the viscosity ratio,
A= ^:.. / rj^. It is seen that break-up will not occur for A > 4 , irrespective of how large
We is. This is because the characteristic deformation time of the drop, given by

''def ~ %^^ext^ % » Ic [8.2.4]

is clearly longer than the time during which the stress acts, t^^^. In the present case,
where G^^^ = rj^^v and r^^^ = l/Vu, it would imply no break-up if A > 1. (This is an
oversimplification, because, e.g., at X = A , [8.2.4] does not hold precisely.) At high A
values, the elongated drop ( D = 5 / 4 / ) starts to rotate in the shear field. At low values
of X, where r^^^ = 1/VD, the drop is elongated into a slender thread. The smaller is
X, the higher is We^^.; i.e., the longer the thread has to become before it breaks. For
X = 10"^ , the draw ratio [L/d] at burst is about 12.

(a ) simple shear flow

T^^x
rr^y o
/
vu = du^ / d y = shear rate
o
1/

(b ) elongational flow
vu = du^ / d x = elongation rate

czzr^-OOOO

Figure 8.11. Two types of two dimensional laminar flow, and the effect on deformation and
break-up of drops at increasing velocity gradient (Vu) . (Redrawn from P. Walstra, Physical
Chemistry of Foods, Marcel Dekker (2002).)
EMULSIONS 8.39

We,

Figure 8.12. Critical Weber number for break-up of drops in various types of flow, as a
function of the viscosity ratio, X . Results (from various sources) of single drop experiments in
two dimensional, simple-shear flow ( a = 0 ), hyperbolic flow { « = ! ) , with intermediate types,
as well as a theoretical result for axi symmetrical elongational flow (ASE). The hatched area
refers to apparent We^^j. values obtained in a colloid mill^^ (Redrawn from Walstra and
Smulders, (1998) \oc. cit.)

As an example of elongational flow, two dimensional hyperbolic flow (curve marked


1 in fig. 8.12) may serve. It is seen that We^^ is much smaller than for slmple-sliear
flow. Qualitatively, this is so because Tr = 2, hence the stress is twice as high, and
also because the stress increases as the drop is elongated. Moreover, the drop does
not rotate, which means that it has enough time to become deformed and be broken
up. A curve for axisymmetrical flow, where Tr = 3 , shows that We^^. is still smaller.
Flow types which are intermediate between simple-shear and plane hyperbolic flow
can also be realized. Some results are in fig. 8.12. The variable is a parameter a in
the velocity gradient tensor, which varies between 0 (simple-shear) and 1 (rotational
component zero). It is seen that a fairly small elongational component in the flow
already lowers We^^ , especially at high A .
Results obtained on (large) single drops under ideal conditions [i.e., very near
equilibrium) agree well with theoretical predictions. This would mean that the value of
d obtained is inversely proportional to / , proportional to Vu and rj, and roughly
proportional to Tr ; it is, further, a function of A, the function depending on the How
type.
However, between the two main drops resulting from break-up, often some much
smaller 'satellite' drops are observed (see, e.g., fig. III. 1.17). These mostly do not form
under conditions where drops of order 1 jim result, since the satellite drops would
then be < 0.1|Lim , implying an excessive Laplace pressure. In other cases, the drops

H. Armbruster, Ph.D. thesis, University of Karlsruhe (1990).


8.40 EMULSIONS

attain a pointed shape with two tips, from which very small drops then emanate. It has
been shown that such 'tip streaming' is due to the presence of a trace of surfactant,
which is swept towards the tips (cf., fig. 8.3b). Also, these satellite drops will generally
not form in practical emulsification.
Assuming the theory to hold under practical conditions, the resulting drops are
predicted to have a diameter that is 2~^^'^ times the diameter corresponding to We^^..
This is, however, not observed in practice, for the following reasons.
- The deformation is too fast to allow equilibrium shapes to establish. It has been
shown in single-drop experiments that the magnitude of dVu/dt can substantially
affect the result. This is especially obvious if dVf/dt is so high that We » ^e^^ can
be reached before break-up occurs. Then a strongly elongated drop results, which
subsequently breaks into many (up to 100) small ones, owing to Rayleigh instability^^
Such 'capillary break-up' appears to occur frequently in practice. The results depend
on the magnitude of rj^l ri^, but it generally implies that the value of d obtained is
smaller than predicted by fig. 8.12.
- Generally, the conditions encountered by a volume element vary among sites in
the apparatus and with time; this applies to the velocity gradient and to the time
during which a drop is subjected to a given stress. Hence, a size distribution of
substantial width results, and it is mostly unclear what kind of averaige should be
taken when comparing theory with results.
- The presence of surfactant can have various effects, depending on the type and
the concentration of the surfactant see section 8.2b.
Some results obtained with a colloid mill are inserted in fig. 8.12. Care was taken
to obtain conditions very close to simple-shear at constant shear rate. The agreement
with theory is reasonable, although break-up also occurred at A > 4 ; presumably,
there was still a small elongational component in the flow.

iii) Turbulent Jlows^\ Turbulent flow is characterized by the presence of eddies


(whorls, vortices), which implies that the local flow velocity, u, generally differs from
its time average value, (u). The velocity fluctuates in a chaotic way and the average
difference between u and (u) equals zero. The root mean square average velocity, u ' ,
is, however, finite and is given by,

ii'^((u-(u))^y/^ [8.2.5]

^^ R.A. de Bruijn, Chem. Eng. Set 48 (1993) 277.


^^ The theory is due to J.M.H. Janssen, H.E.H. Meijer, J. Rheol. 37 (1993) 597. The idea has
been further worked out and experimentally confirmed by J.A. Wieringa, F. van Dieren, J.J.M.
Janssen, and W.G.M. Agterof, Trans. Inst. Chem. Eng. 74 A (1996) 554.
For general aspects, including droplet break-up, see V.G. Levich (1962), and J.T. Davies
(1972), General References.
EMULSIONS 8.41

The value of u' generally depends on direction, but at very high Re (say above
50 000, and small length scale), it hardly does so. The local turbulence then is
isotropic. Kolmogorov has developed theory for this case, which will be used below. It
is largely based on dimensional analysis, which implies that the theory gives scaling
laws. Fortunately, most of the unknown constants in the equations are of order unity.
Turbulent flow shows a spectrum of eddy sizes, C. The largest eddies have the
highest u ' . They transfer their kinetic energy to smaller eddies; these have a smaller
value of u ' , but a higher velocity gradient, u7 C. Small eddies thus have a higher
energy density. They are called 'energy bearing eddies', of length scale C^ . These
transfer their energy to still smaller ones, in which the kinetic energy is finally
dissipated into heat. The smallest eddies are of length i^, also called the Kolmogorov
scale. It is given by

^^=:;;3/4^-l/2^-l/4 [8 2.6]

where e is the power density, i.e., the amount of mechanical energy dissipated per
unit volume and per unit time. Its magnitude ranges from 10"^ to 10^^ Wm"^ for most
emulsifying machines. The variables £ and rj are the most important ones that
determine whether isotropic turbulence occurs, and that characterize its properties.
Local flow velocities depend on the distance scale, x, considered. For a scale
comparable to i^ we have,

u'(x) = fi/3xl/3p-i/3^ X = 0{i^) [8.2.7]

Since the velocity gradient in an eddy is given by u'(x)/x, it follows that the gradient
increases markedly with decreasing size. The characteristic lifetime of an eddy is given
by

T{£^) = i^/uV^) = £l^^€-^^^p^^^ [8.2.8]

For water, the time scale is between 0.1 and lOOjisin most emulsifying apparatus.
The local flow velocity for the smallest eddies is given by

u'(x) = ^l/2x 77-1/2^ x«C^ [8.2.9]

In these eddies, the velocity gradient is therefore independent of distance.


It may be noticed further that [8.2.7] does not contain the viscosity, although this is
a factor that determines the values of Re and C^, and hence the range over which the
equation is valid.

The regime TV will be considered first; here, the diameter of the drop to be broken
up is much smaller than C^ . Break-up can occur if rj^^ is rather high (say, 0.1 Pa s)
8.42 EMULSIONS

and L is not very small {to ensure a high Re value). These conditions will be further
discussed below.
Bi' ak-up of a drop will occur if We > We^^ . Since the flow between two eddies will
always have an elongational component, say a > 0.5 , We^^ will be fairly small and not
depe id greatly on the viscosity ratio A ; see fig. 8.11. The stress acting on the drop
equals r/^Vv. Equation [8.2.9] should be used, because the distance over which the
gradient must act, d« £^, and it yields that Vv = {£/ rj)^^'^ . By using in [8.2.3] for We
the maximum size of a drop that can remain unbroken the result is

d = 2We^^7£:-l/2;y-i/2 ^^^-1/2^-1/2 [8.2.10]

and the size of the drops formed will be given roughly by the same equation. Its
validity has not been checked precisely, to the author's knowledge, but approximate
agreement has been obtained ^^

The regime TI is the common one in the formation of o-w emulsions (and in
making aqueous foams by beating), provided that T]^ is kept small. Now a drop to be
broken up v^ll be of a size comparable to i^ and it will be subject to pressure
fluctuations. From [8.2.7] for u ' , and the Bernoulli equation [8.2.2], it follows that,

Ap{x]^p[u\x)f =£2/3^2/3^1/3 [8.2.11]

If Ap is larger than the Laplace pressure of a drop that is near the eddy, it will be
broken up. Combining [8.2.1 and 11], and equating x with the diameter of the largest
drop that cannot be broken up, we obtain

^^^-2/5^/5^-1/5 [8.2.12]

This equation (actually a scaling relation) is often also used for the average droplet
size obtained, e.g., d^2' This generalization is allowed as long as the size distribution
is of constant shape when, say, the power density is varied. Essentially, the relative
standard deviation C2 should remain constant. This is often the case for high speed
rotor stator stirrers and for high pressure homogenizers, but not always for stirred
vessels. In this last case, a more elaborate treatment is needed^ . The value of C2 (see
[8.1.15]) often is of order unity. It can be reduced by repeating the treatment in the
emulsifying machine, since the various volume elements have not ah passed through
the zones of highest power density in the machine; d^2 is also reduced. See also
fig. 8.17.

^^ See R. Shinnar, J. Fluid Mech., 10 (1961) 259, and especially W.J. Tjabberinga, A. Boon, and
A.K. Chesters, Chem. Eng. Set 48 (1993) 285.
^^ See A.W. Pacek, C.C. Man, and A.W. Nienow, Chem. Eng. Set. 53 (1998) 2005.
EMULSIONS 8.43

Equation [8.2.12] has been very successful in many cases. For a high speed stirrer,
the local value of the power density, i.e., near the stirrer tips, is ^ ~ pco^X'^ where a) =
revolution rate (s~M and X = stirrer diameter. Inserting this in [8.2.12] gives

d32 -X-4/5^-6/5p-3/5j;3/5 [8.2.13]

which is known as the Hinze-Clay relation. The lowest attainable value of d^2 ^^
about 1 |im .
In the valve of a high pressure homogenizer, £ ^ P^^^^ 11, where Phom ^'^ ^^^
pressure drop and t the time of passage of the liquid through the valve slit. The value
of t is inversely proportional to the liquid velocity in the valve, v, which is, in turn,
given by p^^^ ^\p^'^ • Consequently, e - (Phom^^''^ ^^^

d32-Pho4V/^ [8.2.14]

Especially the dependence of droplet size on Phom' ^'^ obeyed very well. Very small
droplets can be obtained, down to d,^^ ^ 01|Ltni.

A condition for [8.2.12] to hold is that the inertial forces acting on a droplet exceed
the viscous forces. Considering that the relevant distance x equals d , [8.2.11] gives
for the inertial stress {u)pVvd, and the viscous stress is given by r/^^v. Hence, the
condition becomes

{u)dp/7]^ = Re^^>l [8.2.15a]

Combination with [8.2.12] yields

d>ril/rp [8.2.15b]

If d = Ti^/Yp, [8.2.10] and [8.2.12] give equal results. The viscous and the inertial
stresses are equal, and we have a regime intermediate between TV and TI.
Another condition for both regimes is that the turbulence is isotropic. However,
even for Re as low as 10 000, the relations seem to hold well. Another aspect is that
turbulence can be depressed by the presence of long polymer molecules in solution,
with contour lengths » ^Q • The result is that the smallest eddies are removed from
the eddy spectrum. Consequently, the resulting value of d^2 ^^ increased and that of
C2 is decreased, at least in regime TI^^ Turbulence depression can also be caused by
the emulsion droplets present if ^ > 0.05. A high value of (p has several conse-
quences, however; see sec. 8.2d.

(iii) Highly viscous drops'^\ It is generally difficult to break-up highly viscous drops

^^ P. Walstra, Chem. Eng. Set 29 (1974) 882.


2)
This aspect is discussed in more detail by P. Walstra, P.E.A. Smulders (1998), General
References.
8.44 EMULSIONS

into smaller ones. We have seen that, in the regime LV, break-up does not occur for a
viscosity ratio A > 4 . For elongational flow, break-up appears possible for any value of
A . However, a general condition for break-up is that r^^^^ (see table 8.2) is not longer
than the time over which the stress acts, r^^^. In practice, elongational flow is
generally induced by sharp constrictions in the vessel through which the emulsion is
forced, and it turns out that T^^^ is often too short for break-up of highly viscous drops
to occur. The result is that the drops obtained are (much) larger than would
correspond with the value of We^^ .
In turbulent flows, the value of r ^ is generally given by the lifetime of an eddy,
^eddv ' ^^^^^ ^h^ value of the eddy size should equal the drop diameter d. In the
regime TV, d« C , and by use of 18.2.9] we obtain Tg^^lv ~ TTC^£~^''^ • Taking r
from table 8.2, it is found that ^^ef/^eddv ~ ^ • ^^ other words, for /I > 1, break-up
would not occur. This is an oversimplification, because drops can also be broken up
by larger, and hence longer living, eddies. The author is unaware of any systematic
theoretical study, but some experimental results are available. In a study with a high
speed stirrer in the regime TV, rf^ was varied, and for lvalues of 0.05, 1.5 and 31,
the resulting 0^32 values were about 3, 4, and 12 |im , respectively^^ Another study
concerned a high pressure homogenizer (regime TI, but low Re ). When rj^ was varied
from 6 to 300 mPa s (at ri^=\], the resulting d^g value ranged from 0.6 to 2.2 |im ;
further increase in 77 had little effect^^
Also in the regime TI, less efficient droplet break-up is to be expected for high X.
The eddy time should be calculated for d^ i^. From [8.2.7] we obtain
^eddv ~ f ^ / ^ p l / ^ / ^ / S pitting this equal to the deformation time as given in table 8.2,
the following equation for the expected value of d at high droplet viscosity results:

d = ^4,,U4pV2 Tdef = r,ddy [8-2-161

Various authors have shown the average diameter to increase with r}^, but the
exponent varied, and was generally < 0.75. Also, the exponent of e varied, generally
being larger than 0.25. Also in this regime further study would be useful. For A values
< 10, [8.2.12] is generally well obeyed, implying that the value of rj^ has very little
effect on the resulting d value.
IVlethods to obtain small drops of high viscosity include the use of surfactants that
give a very small interfacial tension, whereby relatively small d values are obtained at
a given external stress (cf. sec. 8.2c), or by making a w-o emulsion (assuming the oil
phase to be the most viscous one) of ^ == 0.5 , and then inducing a phase-inversion"^.

^' H. Karbstein, Ph.D. thesis, University of Karlsruhe (1994).


^^ W.D. Pandolfe, J. Disp. Set TechnoL 2 (1981) 459.
^^ See e.g. B.W. Brooks, H.N. Richmond and M. Zerfa, in B.P. Binks, Ed., Modern Aspects of
Emulsion Science, chapter 6, p. 175, Royal Soc. Chem. (1998).
EMULSIONS 8.45

(iv) Bounded flow. Droplet break-up in the regime LB will now be considered. In
unbounded flow theories, it is (implicitly) assumed that the velocity gradient is con-
stant over a distance comparable to the droplet diameter. In bounded flow, the
distance between a droplet and the vessel wall, or the next droplet, is 0(d), which
implies that Vv, and thereby the viscous stress, changes significantly over a distance
d . Two situations will be discussed briefly.
The first one is the flow in the valve of a very small high pressure homogenizer.
Here, the valve slit L is very narrow, say < 10 |im . The situation has been analysed by
Kiefer^^ In the absence of particles, Poiseuille flow develops in the slit, the average Vv
being, for example, 10''s~^. Since d = 0(L), the value of Vu over a drop will vary
between 0 and the average Vv. The dynamic thrust of the continuous phase deforms
the drop into a kind of hollow cone, which can be broken up into smaller particles.
The critical drop size would be given by

Mvyp
Some conditions must be fulfilled for the equation to hold, the most important ones
being. Re < 2000 and A > 5 . Further analysis of the flow in a homogenizer valve gives
the result that d is approximately proportional to (Phom^"^'^' ^ relation is
experimentally well obeyed^\ Another difference with the results obtained in a large
homogenizer may be noticed. In the latter, a drop is broken up in several stages,
possibly up to 50 times during one passage through the valve. In a very small valve,
there may be only one break-up event per drop per passage. Hence, quite a v^de drop
size distribution results. To obtain a narrower distribution, with smaller droplets, the
treatment should be repeated, say, five times.
Another situation arises when the volume fraction of droplets is high. Already at
^ = 0 . 5 the average free distance between droplets is of order O.ld . This implies that
the flow around a droplet is markedly bounded. Some interesting results have been
obtained recently. In one study'^^ coarse emulsions of ^ > 0.8 were placed between
parallel plates, one being moved laterally, whereby a virtually constant VD value is
obtained; the slit width L was about five times the droplet size obtained, and Vv
values up to 10-^ s~^ were applied. Almost homodisperse droplets were obtained,
down to 5 |am in diameter. In another study"^^ Couette flow was used [Vv also con-
stant) with larger L and smaller (p (down to 0.4). Here also, quite homodisperse
emulsion resulted, with d from 0.3 to 10 |im , but only if the system, i.e., primarily

^^ P. Kiefer, Ph. D. thesis, University of Karlsruhe (1977). See also P. Walstra (1983), General
References.
^' P.E.A. Smulders, Ph. D. thesis, Wageningen University (2000).
^^ P. Perrin, Langmuir 16 (2000) 4774; P. Perrin, N. Devaud, P. Sergot, and F. Lequeut,
Langmuir 17 (2001) 2656.
"^^ T.G. Mason, J. Bibette, Langmuir 13 (1997) 4600, and C. Mabille, V. Schmitt, Ph. Gorria, F.
Leal Calderon, V. Faye, B. Deminiere, and J. Bibette, Langmuir 16 (2000) 422.
8.46 EMULSIONS

the continuous phase, was markedly visco elastic. The explanation of these effects is
as yet unclear.

8,2c Roles of the surfactant


Generally, emulsions cannot be made without surfactant ^^ The effect of the sur-
factant on the obtained droplet size greatly varies with type and concentration, as
illustrated in fig. 8.13. The plateau values of d^2 ^^e roughly proportional to the
plateau values of j/3/5 ^g predicted by [8.2.12], but the curves vary greatly at lower
surfactant concentrations. For example, the plateau value of d is reached at a lower
concentration for amphiphiles than for polymers, despite the latter generally being
more surface active; see fig. 8.2.
Surfactants play various roles. They facilitate droplet break-up by lowering / , and
thereby the Laplace pressure. They can affect the mode of, and increase the, resistance
to deformation, especially via the surface dilational modulus, K^. In some cases, they
can facilitate droplet break-up by means of surface forces, a phenomenon designated,
'interfacial instability'. Most importantly, they counteract the (re)coalescence of just
formed droplets during emulsifying. All of these phenomena depend on the local, and
often transient, composition of the droplet surface layers.

d32/|im

c / kg m "^

Figure 8.13. Specific droplet surface area A^ and average droplet size d^2 ' ^^ a function of total
surfactant concentration, c , obtained under roughly the same conditions (o/w emulsions,
(p ^ 0.2, regime TI, comparable power density) for various surfactants; PVA = polylvinyl
alcohol). For the soy protein a plateau value is also reached, at c?ii20kgm~^. Approximate
plateau values of y^^ of 3, 10 and 20 mNm~^ for the non-ionic, casein, and PVA, respectively.
Assembled from various sources.

(i) Surfactant transport. This concerns primarily transport of surfactant from


solution (generally the continuous phase) to the drops' surface. Transport is generally
due to convection rather than to diffusion. Table 8.2 gives expressions for the

An example of an exception is making a dilute emulsion of pure oil in pure water by


ultrasonication; the droplets then acquire a substantial negative charge.
EMULSIONS 8.47

characteristic times needed, r^^^ , derived from the flow equations^^ It is interesting to
compare r^^^ with the characteristic deformation time, r^^^, since the ratio of the two
gives an indication of the value of F, and hence of / , during break-up of the drop.
Formulae for r^^^ are also given in table 8.2, derived from [8.2.4]. This equation is
based, however, on the assumption that rj^ » rf^, which Is often not obeyed in the
regimes LV and TV. This is because the deformation time cannot be shorter than the
local value of 1 / Vi;
A few examples will be given for emulsions being formed of oil ( T; = 60 mPa s) and
water (1 mPa s); resulting d = 1 )im ; plateau value of 7"= 2 mg m~^ ; m^ = 2 kg m~^ ;
and at low values of cp , e.g. 0.05. We then obtain for the following cases:

Emulsion type w/o w/o o/w


Regime LV TV TI
'^ads "^ 100 100 1 \iS

^def = 10 10 10 lis

It is seen that for viscous forces r^^^ > r^^^, whereas it is the other way round for
inertial forces, at least in the present cases. The relations are approximate, primarily
because the theory is approximate. Second, it may well be that transport of surfactant
through a very thin layer close to the droplet is by diffusion, further increasing the
adsorption time. The author is unaware of a definitive theory which takes this
phenomenon into account. Probably, any retarding effect will be minor in the regime
TI. Anj^way, adsorption tends to be a fast process. Most polymeric surfactants will
need a time far longer than T^^^ or r^ to attain an equilibrium conformation in the
interface.
During emulsification, the inter facial tension will temporarily vary from place to
place at an interface. Such y- gradients can form because of local variation in adsorp-
tion of surfactant, or because a viscous stress acts at the interface. The latter occurs
frequently during emulsification. If no external force is acting (any more), the
distribution of surfactant at a drop surface will even out. This can occur by three
mechanisms. The first is adsorption of surfactant from the bulk: characteristic times
are given above, and these times depend strongly on the surfactant concentration
(table 8.2). The second one is lateral diffusion in the interface, and the characteristic
time for such a process over a distance z is given by z^ / 4D^ , where the value of the
surface diffusion coefficient, D^ can be approximated by the bulk diffusion coefficient
in the most viscous phase. For a small molecule surfactant, it would be of order
5x10"^^ vcfi s~^. Assuming that d = 1 |j.m , and that z = d / 2 , we obtain a character-
istic diffusion time of somewhat over 10 ms; for a polymeric surfactant, the time will
be substantially longer. The third mechanism is motion of the interface owing to a

See, e.g., Levich (1962), General References, who gives full derivations, or P. Walstra, P.E.A.
Smulders (1998), General References.
8.48 EMULSIONS

7-gradient, which can be described as a longitudinal wave . (This is generally called,


'spreading' of surfactant over the interface, but actually the surface entrains the
surfactant.) The characteristic time to even out a gradient over a distance z is given by

\l/3/A,.\-2/3
T[z) = z^'^{r^pf\^y)-
[8.2.18]
nl/2
/7P = (^wPw) +(^oPo)

Assuming that A/ = 1 mN m~^, which is a value that will often be exceeded, and
z = 0.5 \im, a T(Z) value of about 0.1 ms results. It may be concluded that for small
droplets, evening out of y- gradients is a fast process and that either bulk diffusion or
'spreading' of surfactant over the interface will be the dominant mechanism.

(ii) Transient interfacial tension iy^) • An important variable is the value of y


during drop deformation and break-up, since that determines the final drop size.
Characteristic times for transport towards the drop have been given, but even if the
equations hold, the time needed to reach an equilibrium F value would roughly equal
about 10 X T^^g (for amphiphiles). Moreover, the concentration of surfactant in the
continuous phase, c^., decreases considerably during the process, the decrease in d
causes an increase in A^, the more so for a higher (p, and r^^^ is proportional to
1/CQ. TO give an example, when making an emulsion with ^ = 0 . 3 , resulting in
^32 -^'^ M^^ • this yields A^ = 3 m^ per ml of emulsion. Assuming 7" to become
3 mg m"^, 9 mg of surfactant per ml of emulsion will eventually be on the surface,
corresponding to 1.3% in the continuous phase. If that were the original concen-
tration, and the equilibrium concentration of the surfactant used would be 0.03%, it
would mean a reduction of c^ by a factor of about 40. In practice, surfactant mixtures
are generally applied, which means that the composition of the mixture in solution
also changes during emulsification; see sec. 8.1b. Altogether, break-up would never
occur at the equilibrium value of y.
The last conclusion nearly always holds if there is only one break-up event per drop
during passing of the active zone in a machine. The situation tends to be more
complicated, however, because deformation and break-up of droplets generally occurs
a number of times in succession, by which the surface excess can increase stepwise.
Moreover, when deformation is not followed by break-up, and especially when
coalescence of newly formed drops occurs, the F value is increased. Consequently,
final break-up can occur at a y value that is fairly close to equilibrium, provided that
the initial surfactant concentration is high, although this only applies to amphiphiles
(small molecule surfactants).
This will rarely be the case for polymeric surfactants, because, (a), equilibrium
conformation at the interface is not reached during emulsification and, (b), a higher

^^ J. Lucassen, Trans. Faraday Soc. 64 (1968) 2221.


EMULSIONS 8.49

surface excess (in mass units) is needed to obtain a substantial reduction of y (see fig.
8.2b). Since most polymers also give higher equilibrium values of y than do most
amphiphiles, the droplets obtained will be substantially larger at the same power
density. Consequently, a higher mass concentration is needed to obtain a plateau value
of d32, as illustrated in fig. 8.13. However, this is probably not the only factor
determining the difference in effectivity of both types of surfactant; see subs. (vi).

(Hi) The p.i.t. method. The idea of the phase-inversion temperature and of the p.i.t.
method for making emulsions stems from Shinoda and co-workers . It has been
mentioned briefly in sec. 8.1b.
Consider a system of oil, water (in roughly equal proportions), and 2-7 % of a
surfactant. The latter must be a non-ionic, whose hydrophilic part consists of one or
more poly(oxyethylene) chains, e.g., polyoxyethylene nonylphenyl ether. An idealized
phase diagram, with temperature as the variable, is given in fig. 8.14. At low
temperature, there is an oil phase and an aqueous phase which contains most of the
surfactant. The surfactant is, for the most part, present in micelles, which are swollen
and hence contain oil. Upon increasing the temperature, the micelles become elong-
ated and contain more oil (assuming equilibrium to be reached); moreover, more

Figure 8.14. Idealized phase diagram


of an oil water (non-ionic) surfactant
system that shows a phase-inversion
temperature. For explanation, see
text. (Figure redrawn from Shinoda
and Kunieda(1983).)

phase volume

^' See e.g. K. Shinoda, J. Colloid Interface Set. 24 (1967) 4, and K. Shinoda, and H. Sato, J.
Colloid Interface Set 30 (1969) 258. A review on the p.i.t. and other surfactant properties is by
K. Shinoda, H. Kunieda, in P. Becher, Ed., Encyclopedia of Emulsion Technology, Vol. 1,
chapter 5, p. 337, Marcel Dekker (1983). A thermodynamic interpretation of the p.i.t.
phenomenon is given by E. Ruckenstein, Langmuir 4 (1988) 1318.
8.50 EMULSIONS

surfactant will go to the oil phase. At still higher temperature an apparently homogen-
eous system (designated D) is formed. The interfacial tension between oil and water
becomes very small, and the surfactant still tends to associate. However, the natural
curvature of a monolayer is close to zero (it is concave towards the oil side at low
temperature). The distribution quotient of surfactant between oil and water is of the
order of unity. This is the region near the phase-inversion temperature, and it has
been shown that here the system is a bicontinuous microemulsion^^ Above the p.i.t.
region, the natural monolayer curvature is inverted, and most of the surfactant is in
the oil, in the form of swollen inverted micelles.
The shape of the phase diagram naturally depends on the oil to water ratio. The
phase diagram, including the value of the p.i.t., also depends on the type of oil, the
solvent quality of the water phase for the surfactant (e.g., salt content), and especially
on the concentration and type of surfactant (e.g., length and number of POE chains,
and length and unsaturation of the aliphatic chain).
When the system is agitated in the p.i.t. region, an emulsion forms, of the w/o type
if T > p.i.t., and oAv at T < p.i.t. Shaking by hand suffices to produce droplets down
to d = 1 |im . Close to the p.i.t., the value of y^^ is very small, of the order of
1 0 | i P a s . According to the theories and experiments discussed in sec. 8.2b, an
external stress of order 10 Pa would still be needed to obtain such small droplets, but
such a high value is not nearly reached when shaking by hand. Presumably, droplet
formation is substantially facilitated by interfacicd instability: see subsection (v) below.
In practice, o/w emulsions are often made by the p.i.t. method, because it saves
energy and does not need expensive equipment. On the other hand, the high
concentration of amphiphiles present is not always acceptable and may be costly. The
system is brought to a temperature slightly below the p.i.t., and is agitated with
sufficient intensity. Very small droplets can thus be obtained, even if r]^ is high.
However, they are very unstable towards coalescence, because of the very low
interfacial tension (see section 8.3e). To avoid coalescence, the emulsion has to be
cooled immediately to, say, 20 K below the p.i.t., e.g., by adding cold water if dilution
is acceptable. Naturally, the emulsion should be kept at low temperature. If it is
brought close to the p.i.t., rapid coalescence will occur; and above the p.i.t., true
phase-inversion takes place^^
Water in oil emulsions can also be made by a p.i.t. method, if the p.i.t. is low
enough to make the process feasible^\

^' L. Taisne, B. Cabane, Langmuir 14 (1998) 4744.


Phase-inversion is extensively discussed by B.W. Brooks, H.N. Richmond, and M. Zerfa, in
B.P. Binks, Ed., Modern Aspects of Emulsion Science, chapter 6, p. 175, Royal Soc. Chem.
(1998).
^^ H. Sato, K. Shinoda, J. Colloid Interface Set 32 (1970) 647.
EMULSIONS 8.51

(iv) Effect on drop deformation. The main effect of the surfactant on droplet
deformation and break-up, viz., the lowering of y, has been discussed above. There
are some other effects related to the / gradients (see section 8.1c) that are formed by
flow of the continuous phase along the drop. In principle, the presence of such a
gradient can stop internal circulation in the drop, which would lower the energy
needed for break-up. Moreover, the variation of y over the surface of a drop affects
the deformation of the drop. In a largely theoretical study in the regime LV, the overall
effect on break-up appeared to be small to moderate^\ However, when tip streaming
occurs (see section 8.2b sub (i)) in the regime LV, and the stresses are such that
relatively large drops will result, the formation of small satellite droplets appears to be
possible, since bimodal size distributions were obtained experimentally. However, in
practice it took a very long emulsification time to arrive at a substantial volume
fraction of small droplets^^
Another effect is that the Gibbs energy needed for deformation of a drop equals the
increase in Ay; generally, both A and y increase during deformation. Since d{Ay) =
ydA + Ady, and the surface dilational modulus K^ = dy/d\nA, the following relation
results

d{Ay) = {y+K^)dA [8.2.19]

It would thus be OS if 7 is enlarged by the value that Kg attains under the prevailing
conditions. The situation has been studied in the regime LV on single drops, and

>^ observed

0.5 k predicted ^ s .
Figure 8.15. Effect of sur-
factant concentration, c^-., on
break-up of a single drop in
simple-shear flow. The value x
is the critical droplet size for
break-up relative to that in the
2hk absence of surfactant, as ob-
served and as calculated for the
equilibrium value of y. The
quantity y is the critical Weber
number obtained over that pre-
dicted for 7eq • (After results by
1 1 1 1 1 Janssen et al. (1994) loc. ctt.)
101-3 10-^ 10
Cp/mol m~^

^^ H.A. Stone, L.G. Leal, J. Fluid Mech. 64 (1990) 161, and W.J. Milliken, L.G. Leal, J. Colloid
Interface Set 166 (1994) 275.
^^ F. Groeneweg, F. van Dieren, and W.G.M. Agterof, Coll. Surf. A91 (1994) 207.
8.52 EMULSIONS

under conditions where y^^ was close to its equilibrium value at the prevailing
surfactant concentration^^ The main result is given in fig. 8.15. It can be interpreted
as if the interfacial tension during break-up equalled y + hK^, the value of the factor
h being 0.23 in the case studied; it is uncertain why h < 1 (because flow in the drop is
inhibited?). It is seen that the effect can nevertheless be substantial. To what extent
this is also true in other regimes, or for polymeric, surfactants during practical emul-
sification is still unclear.

(v) Interfacial instability^\ This phenomenon is also called 'interfacial turbulence'


and incorrectly 'spontaneous emulsification'.
Consider two liquid phases (1 and 2) that are separated by an interface having a
very small interfacial tension. Surfactant will diffuse from phase 1 to phase 2 if its
chemical potential is higher in 1 than in 2. This means, in practice, that a surfactant is
chosen that is soluble in both phases, but (far) more so in phase 2. It is dissolved in
phase 1, and then the phases are brought into contact. Now an unstable situation
arises. A small deformation of the interface will be enlarged if the bulge is toward
phase 2. Because of depletion of surfactant in the bulge, y gradients are formed, and
these cause Marangoni flow, as illustrated in fig. 8.16. Consequently, fairly long fingers
of phase 1 can protrude into phase 2, and are then subject to Rayleigh instability,
breaking into small drops. The extent to which interfacial instability occurs depends
on some factors; a significant effect appears to be more likely if 7/^ > 7/2 . Moreover, the
electrostatic charge of the interface, and the thickness of the electric double layer have
a substantial effect^\
It is presumably a combination of gentle agitation (implying fairly long deformation
times) and the generation of interfacial turbulence that can result in the formation of
quite small drops. Prerequisites concerning the surfactant are (a), that it causes a very
small y^^ (< 0.1 mPa s), Implying a suitable (amphiphilic) surfactant at high concen-
tration; (b), that it is well soluble in the disperse phase, as such, or in micelles; and,
(c), that it is nevertheless more soluble in the continuous phase. Especially for the

phase 1
Figure 8.16. Illustration of an interfacial
phase 2 \ , \ \ X ^ small instability. Phas(
Phase 1 originally contains the
surfactant whicl
which is, however, more sol-
uble in phase 2.

/ large

^^ J.J.M. Janssen, A. Boon, and W.G.M. Agterof, Coll. Surf. A91 (1994) 141, and A.I.Ch.E.J. 40
(1994) 1929.
^^ This phenomenon has been studied by C.V. Sternling, L.E. Scriven, A.I.Ch.E.J. 5 (1959) 514,
and by J.H. Gouda, P. Joos, Chem. Eng. Set 46 (1964) 521.
Theory has been developed and checked by A. Sanfeld, M. Lin, A. Bois, I. Panaiotov, and J.F.
Baret, Adv. Colloid Interface Sci. 20 (1984) 101.
EMULSIONS 8.53

making of emulsions with a very viscous disperse phase, e.g., of a bituminous oil in
water, application of these principles may be useful. As mentioned, inter facial instab-
ility will also be of importance when applying the p.i.t. method of emulsification.

(vi) Prevention of re-coalescence. If emulsification occurs at a given intensity, i.e., at a


given value of €, and if the intensity is then changed to a lower value while
emulsification proceeds, the average drop size increases in time, until a plateau value
is reached. This has been observed by several workers, at least in the regimes TV and
TI. It must be caused by coalescence of newly formed drops, and the phenomenon is
generally called re-coalescence. For example, in systematic studies using a small high
pressure homogenizer^\ the coalescence rate was higher for a higher value of (p
although the rate constant for coalescence (^2, see below) was smaller for higher cp.
The value of ^2 was higher for a larger initial droplet diameter, a higher £ value, and
a lower surfactant concentration, hence presumably a lower F value.
These results agree well with the occurrence of coalescence owing to orthokinetic
aggregation (see sec. IV.4.5b). It is generally assumed that, during emulsification,
break-up and coalescence occur simultaneously, as depicted in fig. 8.10, and that
eventually a steady state is reached. This is illustrated in fig. 8.17, where it is seen that
both the average and the standard deviation of d go to a plateau value. A simple equa-
tion is sometimes suggested, viz., diV^/dt = kjiV^ - k 2 N ^ , where N^ is droplet num-
ber concentration and k^ is a break-up rate constant. The final droplet number would
then be given by fc^//c2. However, this equation does not fit the results at all^\ be-
cause both rate constants depend greatly on droplet size and other changing variables.

2
(a) (b) J
\^10/(^10)co
\^2/(C2)oo

1
^^32/^32)00

Ultra Turrax homogenizer

1 1 1 1 1 1 1 1
10 20
t(ks) number of passes

Figure 8.17. Average droplet size, d, and size distribution width, c, relative to their plateau
values, as a function of agitation time or number of passes. In (a), the surfactant was a
commercial amphiphile; in (b), a protein mixture. (Adapted from Walstra, (1983), ioc. cit.)

^'S. Mohan, G. Narsimham, J. Colloid Interface Set 192 (1997) 1; and G. Narsimham, P.
Goel, J. Colloid Interface Set 238 (2001) 420.
^^ See P. Walstra (1983), General References.
8.54 EMULSIONS

Other methods for estimating the rate of re-coalescence depend on establishing the
extent of mixing of oil from different droplets during emulsification, for example when
droplets having two values of the refractive index, n , are present. To carry this out,
two emulsions are made under identical conditions, but the (oil) drops differ in n
value (Hj and 1X2 ). The emulsions are mixed and the mixture is emulsified again at
the same intensity, during various times or with an increasing number of passes. The
emulsions obtained are diluted with an (aqueous) solution of high refractive index so
that the continuous phase attains a value of n = [11^ + 11.2) • The turbidity of the diluted
mixtures is determined. If no mixing of oil, i.e., no re-coalescence, has occurred, the
turbidity will equal that of the (dilute) original mixture; if re-coalescence would be
complete, the turbidity will be virtually zero owing to refractive index matching. This
allows estimation of the extent of re-coalescence ^^ An alternative is to make two
disperse phases, one of which contains a fluorescent probe plus an excimer, at such
concentrations that dilution with pure oil virtually eliminates the fluorescent signal^^
The advantage of these methods is that the measurement of re-coalescence occurs
under nearly the same conditions as the emulsification. Moreover, comparison of the
variation in the average drop size obtained by, for example, varying surfactant type or
concentration can be correlated with variation in the re-coalescence rate. Roughly the
same trends are observed as with the other method; examples will be given below.
Coalescence can occur if two droplets collide (a phenomenon that will occur
frequently during emulsification): if and when the film between them becomes very
thin, say a few nm, it will generally rupture, inducing coalescence. It is often assumed
that prevention of close approach, and hence of re-coalescence, is due to colloidal
forces, i.e., electrostatic or steric repulsion, caused by the presence of surfactant at
the droplet surface. This can be questioned. The stress which presses droplets
together during emulsification can be quite large, up to the value needed to result in
droplets of that size (say, half the Laplace pressure of the drops). Considering a drop
of d = 1 |Lim and with 7 = 0.01Nm~^, the maximum stress will be 20 kPa. From
DLVO theory, it can be calculated that the disjoining pressure between the droplets
will, under most conditions, be < 200 Pa, even for drops with a plateau value of F.
For steric repulsion, the values may be higher, but generally not much. Hence, the
repulsive forces are generally far too weak to prevent re-coalescence.
This is borne out by the following experimental result"^I An o/w emulsion was made
with SDS under such conditions that insufficient SDS was present to give full
monolayers (r==0.06 F^]. Nevertheless, coalescence was not observed within a day.
Adding 10 mmolar NaCl, which greatly reduces electrostatic repulsion, led to visible
coalescence. When the same emulsion was made in the presence of 10 mmolar salt, it

^^ L. Taisne, P. Walstra, and B. Cabane, J. Colloid Interface Sci. 184 (1996) 378.
^' L. Lobo, in Proc. 2"^ World Congress on Emulsion, Vol. 4 (1977) 75.
^^ L. Taisne et al., (1996) loc. cit.
EMULSIONS 8.55

nevertheless gave exactly the same value of d^2 ^^ without the salt, i.e., 0.24 |im (to
be sure, this emulsion also showed coalescence upon standing).

To find out what mechanisms are responsible for diminishing re-coalescence,


kinetic aspects should be taken into account: what happens with the droplets as a
function of their surface excess during agitation? A full treatment is elaborate ^ ^ and we
will mention only some trends, since accurate prediction is not possible, anyway,
because of the complex and constantly changing situation during emulsification^^ It is
assumed that the surfactant is soluble in the continuous phase only.
An important question concerns the transient value of F during a collision of
drops. This will depend on the ratio of the characteristic adsorption time divided by
the average time between encounters ('collisions') of a drop with another drop, r^^^^ .
From hydrodynamic theory it follows that in all regimes in unbounded flow, the
(average) ratio is given by,

^ads/W=50rp,^/Ccd 18.2.201

Assuming F^ /c^ =1 \im , cp = 0.03 , and d = 1 |am , the ratio is 1.5. In most cases,
e.g., with larger (p or smaller c^., the result will be that r^^^g is longer, or even much
longer, than r^^^ . It then means that F is often substantially smaller than r j during
a droplet encounter. On the other hand, when a drop starts to deform, it may already
have acquired a substantial F value, and if the total surfactant concentration is high
enough, many drops may even have F ~ r j . As soon as a drop is deformed or even
broken up, however, the value of F can become substantially decreased; nevertheless,
adsorption does not start with a bare interface. Considering further that considerable
statistical variation occurs, particularly in the value of r^^^ , there will be frequent
occasions where F is low during an encounter.
Owing to the large stress acting on them, the drops will often be deformed upon
approach, forming a flat film that can subsequently drain. This is illustrated in fig.
8.18. The rate of drainage is an important variable.
We will first consider a situation where the droplet surface is rigid in the lateral
direction owing to the / gradient formed. The Reynolds equation can be applied, and
for the present situation the linear drainage rate is given by

dh -2h3
Sny^h Syh^
v=- [8.2.21]
dt Sri^a^F 3 7]^ a^

The following discussion is for the most part based on A.K. Chesters, Trans. Inst. Chem.
Eng. E69,B (1991) 259; and on LB. Ivanov, P.A. Kralchevsky, Coll. Surf. A128 (1997) 155. Also
see Chesters (1991) and Ivanov and Kralchevsky (1997) in the General References.
Discussed by Walstra and Smulders (1998), General References.
8.56 EMULSIONS

(1) (2) (3)

Figure 8.18. Subsequent stages in the formation and drainage of a film between two drops that
approach each other owing to a high stress. In frame (1), the motion of drops and continuous
phase are indicated; in (2) is shown the distribution of surfactant molecules (indicated by
dashes); and (3) defines the geometry of the system. Highly schematic and not to scale.

It is seen that a larger force (F) gives slower drainage, and hence a longer, not a
shorter, drainage time. This is due to the higher stress causing a flat film to form of
larger radius, R, and hence a greater resistance to flow. It is indeed frequently
observed, though not always, that more intense agitation, and hence a higher local
stress, goes along with less coalescence.
The quantity between parentheses in [8.2.21] is valid for F = nya, which derives
from the assumption that the stress acting on the drops equals half their Laplace
pressure. Integration yields for the time needed to drain the film to a thickness h, the
relation

Mr [8.2.22]
16r/i2

Again, let d = 1 |j,m , T]^ =1 m P a s , and ;^ = 10 mN m"^, the result is for drainage to 5
nm, r^j. = 90 jis ; for d = 3 jim , the value will be 2.5 ms. These times appeair quite
short, but the time during which a high stress will act on a droplet pair is even
shorter. For the regime TI, [8.2.8] yields, assuming that ^^ = lOd , a lifetime of 1.5 jus
for d = 1, a value of 8 jis results for d = 3 jum. For the regime TV, roughly the same
values are obtained. This means that for the present case, i.e., a rigid drop surface,
the time during which the drops are strongly pressed together is generally too small to
obtain an h value small enough for film rupture to occur. In other words, the drops
move away from each other before they can coalesce, at least in turbulent flow. (This is
probably also true in elongational flow, but not always in simple-shear flow.)

For non-rigid surfaces the situation is different. The extreme is a surface that is
fully mobile In the lateral direction. The drainage is then given by
EMULSIONS 8.57

h =hoexp(-t/^^jJ
^ch^^c^^y regime TV [8.2.23]
t^^^pd^{e/7]^f^'^ /lOy regime TI

where h^ is the original distance between the drop surfaces, and t^^ is the char-
acteristic drainage time. Assuming h/1x^ = 0.05 (drainage from h = 100to 5 nm),
d = 1 )j.m , 7 = 30 mN m~^ (since a fully mobile surface implies F close to zero) and
7 7 ^ = l m P a s , we obtain for r^^ in the regimes TV and TI, 0.1 and 0.05 |j.s , res-
pectively; for d = 3 |im the results are about 0.3 |is in both regimes. This implies that
the drainage times can be substantially shorter than the duration of the local stress,
and hence that coalescence would occur readily, which does indeed happen in the
absence of surfactant.
In nearly all practical situations, the surface will be 'partially mobile', which implies
that it does move under the prevalent tangential stress, but more slowly than in the
absence of surfactant. The important variable is the Marangoni number; for a rigid
surface Ma > 1 is assumed, for a fully mobile surface Ma ~ 0 . From [8.1.4] Ma can be
calculated. In the present case, the stress acting in the film is about y/a , which leads
to Ma ~ K^/ y • Equation [8.1.5] gives an expression for K^ and, in the present case,
the value of the parameter ^ will generally be negligible, owing to the very short time
scale. This implies that the surfactant forms virtually a Langmuir monolayer. The
result is
dn d In 7
Ma. = = ^ [8.2.24]
^"^ ydlnr dlnr
Calculation of the drainage times as a function of Ma is complicated, and there is
some disagreement among authors. However, the drainage time is probably not much
shorter than for a rigid surface (see [8.2.22]) if Ma^^ > 0.5. Equations of state, as
given, for example in fig. 8.2.b, allow calculation. A value of Ma^^ > 0.5 will be
reached for SDS at F ^ 0.4 , and for p-casein at about 1.1 mg m~^ . The latter value is
probably too low: at the very short timescales involved during emulsification, the K
values at a given low F must be even smaller than the equilibrium values given in the
figure for this flexible protein.
As mentioned earlier, a considerable difference is observed between the d^2 values
obtained with amphiphiles and with polymers at the same intensity of agitation, and
the same mass concentration of surfactant; the difference is especially large at a
relatively low surfactant concentration. The above considerations point to the need for
a sufficiently high Ma value during drainage of a film between droplets to prevent their
re-coalescence. For polymers this is more difficult to achieve, they require a much
higher surface concentration (in units of mass per unit of surface area) to attain a
substantial Ma value. The latter will be due primarily to the large molar mass of
polymers.
8.58 EMULSIONS

The importance of molar mass is further illustrated in the following case study.
Two surfactants were used. The one was P- casein, a flexible protein of 209 amino acid
residues, with part of the peptide chain being highly charged, and another, longer part
being less charged and having a high concentration of hydrophobic side groups. The
other was a fragment of the same (3- casein, obtained by selective hydrolysis, consisting
of residues 29-106, and hence a much shorter chain; nevertheless, it has a similar
amphiphilic character. Both of these surfactants were used to make emulsions, with
the results shown in fig. 8.19. It is seen that the peptide gave smaller d^2 values at the
same surfactant concentration. The rate of re-coalescence was estimated by the
refractive index matching method, from the relative decrease in turbidity during
homogenization (because a very small homogenizer was used, a great number of
homogenization steps is needed to obtadn a steady state). Figure 8.19b shows that the
rate of re-coalescence is well correlated with the difference in d^2 • Figure 8.19c shows

•xj

0.5

5 10 20 40
c/mg ml" ^ passes

Figure 8.19. Emulsions made with p-casein (BC) and an amphiphilic peptide (AP); see text, (a)
The resulting values of d^2 ^^ ^ function of total surfactant concentration c. (b) Relative
turbidity as a function of the number of passes through the homogenizer valve, (c) Values of
dr^2 ^s a function of storage time t of emulsions made at c — 2 mg/ml. Emulsions of ^ = 0.2 ,
pH = 6.7, ionic strength = 75 mmolar .

After P.E.A. Smulders, P.W.J.R. Caessens, and P. Walstra, in E. Dickinson, J.M. Rodriguez
Patino, Eds., Food Emulsions and Foams, Royal Soc. Chem. (1999) p. 61.
EMULSIONS 8.59

that the difference is inversely correlated with a difference in coalescence stability of


the finished emulsions. It is also unlikely that the peptide gave a lower value of y^^
thain the protein.
Further studies on re-coalescence^^ have shown that it occurred less if the emulsion
was re-homogenized at a lower value of e. For SDS as the surfactant, the re-coales-
cence rate was very small if the surfactant was in excess. Good correlations were
obtained between re-coalescence rate and the droplet size obtained when varying the
surfactant concentration, using various peptides and proteins, and upon varying the
pH. However, when comparing the proteins P-lactoglobulin and (3-casein, the former
gave droplets that were smaller by some 25%, whereas the re-coalescence rates
appeared to be the same.
In summary, much of the differences given between various surfactants and
surfactant concentrations can be ascribed to differences in re-coalescence, and the
mechanisms involved are understood, at least in a qualitative sense. The issue is
somewhat confused by the other effects a surfactant can have. To arrive at hard
conclusions, more systematic studies would be needed, taking into account the
considerations given above.

(vii) BancrqfVs rule. The reasoning in the above subsection is based on the
assumption that there is one surfactant that is soluble in the continuous phase only .
If the surfactant is in the drops, and is not transferred to the continuous phase, y-
gradients hardly form, since surfactant can readily diffuse to the thinnest spot in the
film between approaching droplets. It has been shown'^^ that drainage will be prac-
ticEilly along a mobile surface; in other words, Ma^^ will be close to zero in nearly all
cases. This provides directly an explanation for Bancroft's rule, when making an emul-
sion of two immiscible liquids and a surfactant, the liquid in which the surfactant is
better soluble will become the continuous phase, since droplets containing surfactant
will recoalesce much faster than those without dissolved surfactant.
It should be added, however, that the reasoning given in subsection (vi) does not
apply to the situation described in subsection (iii), where the p.i.t. method is des-
cribed. Here, the time scales involved are much longer, the forces involved are very
much smaller, and the interfacial stability described in subsection (v) plays an import-
ant role. A different treatment is needed, also, to explaiin Bancroft's rule^\

8.2d Formation of surface layers


When making an emulsion with an amphiphilic surfactant (mixture), the surface

^'See L. Taisne, P. Walstra and B. Cabane, J. Colloid Interface Set 184 (1996) 378, for
emulsification with SDS, and P.E.A. Smulders, Ph.D. Thesis, Wageningen University, (2000), for
proteins and peptides.
See I.B. Ivanov and P.A. Kralchevsky, mentioned in the General References.
^^ See e.g. A. Kabalnov, H. Wennerstrom, Langmuir 12 (1996) 276.
8.60 EMULSIONS

excess and the composition of the adsorption layer will be governed by the Gibbs
equation [8.1.2], or equations elaborated for more complicated systems. In principle,
r can be calculated from the values of (p, the total surfactant concentration c , and
d^2 ^^ \ • T'his is generally not so for polymeric surfactants.
The main point is that equilibrium is generally not reached. Figure 8.20a shows an
adsorption isotherm obtained for (3-casein (a flexible, almost random coil protein in
solution) on a macroscopic surface, alloMring sufficient time to obtain equilibrium. It
also gives an apparent adsorption isotherm, calculated from F and c^ values determ-
ined after emulsification, for a range of initial surfactant concentrations. The differ-
ences are striking. The j u m p in the 'macroscopic' curve around c^^ = 10^ is unex-
plained (the increase in F is real, but it is not known precisely what the shape of the
curve is). It appears that this increase in F does not correspond to a significant
decrease in / . Neither is it known whether the y value on the emulsion droplets with
F ~ 4r mg m~^ is lower than that observed in the macroscopic experiment ( TT ~
20 mN m~^). If an emulsion has been made under conditions such that a relatively
small F value results, e.g. , 2 mg m~^ , addition of p-casein to the emulsion causes an
increase in F. The, 'emulsion' curve also depends on other factors, such as the value
of (p during emulsification. If the data are plotted as in fig. 8.20b, the same curve is
observed for different cp values. This means that the magnitude of F depends not
only on c^ , but also on the amount of continuous phase per unit droplet surface area.
Figure 8.20b also shows that different proteins give different relations. By and
large, for proteins that do not differ greatly in other properties, there is a correlation

/ 1 y^
/ //BC
s if
6 J y"^ -

'Ix
2 \-
f emulsion f 1/ BL , LY

1 h /macroscopic ^ /
/ interface (a) / (b)
/
Ln 1 1 1 1 1
10- 10^ 10^ 0 10 15
C -2
c^/ppm — m g m -^
Ay
Figure 8.20. Values of the surface excess, F, at the triglyceride oil water interface for some
proteins, (a) p- casein; F is plotted against protein concentration in the sub natant (left-hand
curve) and in the continuous phase of an emulsion (right-hand curve); see text, (b) F plotted
against total surfactant concentration over the surface area produced for various proteins, viz.,
p- casein (BC), ovalbumin (OV), p- lactoglobulin (BL) and lysozyme (LY). The broken line
indicates what the result would be if all of the protein were adsorbed .

' For the most part, after P.E.A. Smulders, Ph.D. Thesis, Wageningen University (2000).
EMULSIONS 8.61

between molar mass and the plateau surface excess. The Increase In F at high c/ A^
values appears to be due to aggregation of the protein in the interface. (This cannot
explain the differences in fig. 8.20a: P-casein does not show aggregation.)

8,2e Effect of volume fraction


Figure 8.21a shows d32 as a function of homogenization pressure for various
values of cp. The initial concentration of surfactant in the liquid making the
continuous phase was constant. The regime was Tl. For small cp values, [8.2.14] was
obeyed, the slope of logd32 versus logpj^ being exactly 0.6. The higher is (p, the
larger the values of d^2 became, the more so at a higher homogenizing pressure.
To explain the observations given in fig. 8.21, possible consequences of increasing
the value of (p will be considered.
a. The encounter frequency of a drop will increase; for small (p, in proportion to cp .
This will enhance the probability of re-coalescence and may thus lead to a larger drop
size.
b. The value of F^^^ will decrease, unless there is a large excess of surfactant. The
depletion of surfactant from the continuous phase will be stronger, since the value of
A^ will increase more. This will, in turn, cause the value of y^^^ to be higher during
break-up, causing a larger average d . Moreover, the value of the Marangoni number in
the thin film between approaching drops may well be smaller, leading to an enhanced
probability of re-coalescence during an encounter, and hence a larger average d . See
sec. 8.2c sub (vi).
c. The viscosity of the emulsion will increase, which generally means a decrease in
Reynolds number, which may lead to a change in regime, e.g., from Tl to TV.

-^^^....,,,0^
0
•c"
^^^Vv<C).3
O
0.2
^ O N S O . 2

0.4
(a) ^0.05\.
1 1 1 iS
0.6 1 1.4
log (Ph/MPa)
(P
Figure 8.21. Effects of droplet volume fraction, cp, on the resulting droplet size in
emulsification. (a) Triglyceride oil in water emulsions made in a high pressure homogenizer at
various pressures p^. The surfactant is milk protein. (After results by P. Walstra and G. Hof,
unpublished.) (b) Toluene in water emulsions made by stirring in a small vessel at various
values of (p. The dj^^x ^^ ^^^ largest drop size observed in the emulsion by microscopy. The
surfactant was SDS, present at high concentration. (Redrawn from results by S. Kumar, R.
Kumar, and K.S. Gandhi, Chem. Eng. Sci. 46 (1991) 2483.)
8.62 EMULSIONS

d. Particles depress turbulence, the more so at a higher concentration (see section


8.2b sub (ii)); in particular, the smallest eddies will be removed from the spectrum. In
the regime TI this will lead to an increase in drop size.
e. The effective viscosity of the continuous phase, r}^, will increase. To be sure, its
microscopic viscosity will remain the same, but the drops will sense a higher viscous
stress. For the most part, the drops will have a tangentially rigid surface during
agitation. This means that the velocity gradient between two drops will be enhanced at
the same agitation intensity, as if rj^ is increased. Moreover, the flow will have a
stronger elongational component, which also means a higher viscous stress. If the
drops are broken up by viscous forces, this may cause a greater reduction in average
drop size.
The results of fig. 8.21a can be explained by factors (a), (b), and possibly (d). The
regime was probably TI in nearly all cases.
The results of fig. 8.21b are more complex. Here, the power density was quite low;
at low values of cp, the regime was undoubtedly TI. Since the surfactant was in excess,
factors (a) and (b) probably had little effect (especially since SDS is quite effective at
fairly low concentrations). The increase in d must then, for the most part, have been
due to increasing turbulence depression (d). Then a change in regime may have
occurred (factor c), but the authors have a slightly different explanation. They argue
convincingly that the various mechanisms for break-up all act, in all cases, but that
conditions determine which of these mechanisms is the most effective and thus
determines the droplet size obtained. Below ^ = 0.45, that will be inertial forces,
above ^ = 0.5, viscous forces take over. The further decrease in d^^ with increasing
(p v^ll then be due to factor (e), i.e., an increase in effective 7]^ .
In other cases, other relations may be found, but the same variables should be
considered. Quantitative prediction may be difficult.

8.3 Stability

Emulsions are thermodynamically unstable, because they have an excess of interfacial


Gibbs energy; this results in a tendency to decrease the specific interfacial area A^ .
We consider here changes in dispersity, i.e., changes in size, shape and position of the
droplets. The processes involved may, however, be very slow, because the system can
be in a metastable or a frozen equilibrium state (sec. 1.2.3). These aspects will be
treated in this section, which first gives an overview of the various instabilities and
then discusses each of these in more detail ^\
In most cases, stability is desired: changes in dispersity should be insignificant
during the intended lifetime of the system. In other cases, instability is desired, for
example to achieve recovery of the disperse phase, or when making an emulsion gel.

We will largely follow an earlier review (P. Walstra, 1993) and chapters 1, 6, 7 and 9 in B.P.
Sinks (1998); sec General references.
EMULSIONS 8.63

8,3a Overview
Figure 8.22 depicts the most important physical instabilities that can occur in
emulsions.
Ostwald ripening. Material in small drops can diffuse through the continuous
phase toward larger drops. This leads to an increase in the average droplet size. The
driving force is the difference in chemical potential of the material, caused by the
difference in curvature, as expressed in the Kelvin equation. A prerequisite is a finite
solubility of the disperse in the continuous phase.
Aggregation. When droplets stay together for an appreciable time after having
encountered each other as a result, for example, of Brownian motion, they are said to
be aggregated. The aggregates can grow to considerable size, and possibly form a
space filling network. The driving force is often van der Waals attraction between
drops, or depletion interaction caused by polymer molecules or surfactant micelles
present in the continuous phase; moreover, some kinds of bridging of drops can cause
aggregation. Aggregation can be counteracted by colloidal repulsion. Hence, deaggreg-
ation can often be achieved by changing the disperse phase composition.
Sedimentation of drops is caused by gravitational forces acting on the drops; it
results in a decreased potential energy of the system. Sedimentation can be upwards,
i.e., creaming, or downwards, i.e., settling. The final result is separation of the system
into a cream layer (or sediment) of high (p, and a layer of continuous phase. Sedi-
mentation can generally be undone by mild agitation.
Coalescence occurs when the thin film between close drops ruptures; the local
differences in Laplace pressure then result in the merging of the two drops into a
larger one. The driving force is the decrease in interfacial area that results from
coalescence. Ongoing coalescence results in complete separation of the phases (into
two layers).

Ostwald ripening
o-o-o
O o Figure 8.22. Illustration
of changes in dispersity of
aggregation o o QD ft) an emulsion. Broken ar-
rows denote the reverse
change. Ostwald ripening
and coalescence (actually,
the second arrow in the
lowest row) are irrevers-
sedimentation
O TT-^' ible.

coalescence
oo-co-O
8.64 EMULSIONS

The instabilities mentioned should be clearly distinguished, since they are governed
by different variables and have different consequences. It is not always easy to dis-
tinguish them experimentally, especially in an early stage. Moreover, the processes can
affect each other. All changes that cause an increase in particle size will enhance the
sedimentation rate. This holds especially for aggregation, since its rate is, in turn,
enhanced by sedimentation. Coalescence can only occur if two drops are quite close to
each other. Consequently, it is enhanced in aggregates and in a cream layer or
sediment.

The decrease in Gibbs energy involved in the changes mentioned tends to be


small. Consider an o/w emulsion of (p = 0.l, d = l }im and y^^ =10 mN m"^. If com-
plete coalescence occurs, the decrease in Gibbs energy will equal 6 7 ^ / d = 6 k J m"*^. If
the system is isolated, the release of this energy would increase its temperature by a
mere 1.5 mK. For the other instabilities, the decrease in Gibbs energy will be even
smaller. However, the slowness often observed for these processes is not due to a
small driving force, but either to a substantial activation Gibbs energy (hence a
metastable state), or to slowing down Brownian or other translational motion of the
drops. Aggregation, sedimentation, and coalescence cannot occur if the particles are
somehow immobilized. In practice, most emulsions tend to be more stable if the
droplets are smaller, except for Ostwald ripening.

Another type of change in dispersity is called phase-inversion, i.e., the change of an


o-w emulsion into w-o, or vice versa. Two types can be distinguished. Transitional
phase-inversion can occur near the phase-inversion temperature (p.i.t.) of an emulsion
made with a poly (oxy ethylene) containing surfactant, where the inter facial tension is
virtually zero. This has been discussed in sec. 8.2c, sub (iii). The other type is
catastrophic phase-inversion, which can occur in some systems if the volume fraction
is made very high, owing to the addition of disperse or the removal of continuous
phase. This inversion is a complex phenomenon. It is affected by the nature and the
concentration of surfactants, the intensity of agitation, and the order in which
substances are added or removed. The process is poorly understood and will not be
discussed further . Incidentally, it is often possible to reach volume fractions up to
about 0.98 without phase-inversion occurring, provided that the Laplace pressure of
the droplets is rather small.

Finally, the presence of a third phase may affect stability. Air bubbles beaten into
an o/w emulsion may cause spreading of oil over the a-w interface, and cause dis-
ruption of the drops into smaller ones. Surfactants that can give rise to lamellar liquid

See chapter 6 in B.P. Binks (1998), General references, for an extensive review on phase-
inversion.
EMULSIONS 8.65

crystalline phases with water, may, if present in a sufficiently high concentration,


form such phases around oil droplets, and thereby enhance the stability against
coalescence ^\ Small solid particles present in the continuous phase that adsorb onto
the drops, can counteract coalescence. This 'Pickering stabilization' will be discussed
in sec. 8.4. Particles present in the drops, especially triacylglycerol crystals, can in
principle cause gross instability owing to 'partial coalescence'; this is discussed briefly
in sec. 8.3c.

8.3b Ostwald ripening^^


The solubility of the material in a small particle is generally increased over the bulk
solubility c^^^ ^ . The increase in solubility is given by the so called Kelvin equation
(although the form in which it is used is, in fact, due to Ostwald); see sees. 1.2.23c and
IV.2.2e. For a spherical particle of radius a the equation reads;

Sat^«^ = ^sat.ooexp(x7a)
[8.3.1]
X' = 2YM^/P^RT

where M^ and pj^ are the molar mass and the density of the material in the disperse
phase, respectively. The parameter x' is a characteristic scale, in units of length; the
larger it is the greater the value of c^^t^^^/^satoo • Consider a dodecane droplet in
water. The interfacial tension is about 0.05 N m~^, Mj^ = 0.17 kg mol~^, and
Pj) = 745 kg m~^; at room temperature this results in x' ~ 10 nm . For a droplet of
a = 0.5 |im this results in a solubility ratio, Cg^^(a)/Cg^^^ = 1.02 , i.e., significantly
greater than unity.
If the droplets in an emulsion are polydisperse, and the disperse phase has a finite
solubility in the continuous phase (as is the case for dodecane in water), this will
result in Ostwald ripening: the smaller drops decrease in size and eventually
disappear, whereas the larger drops grow. The driving force is the difference in
chemical potential, and hence in solubility, of the material between small and large
drops. The material is transported by diffusion; the mass transfer process can be
called isothermal distillation; and the result is disproportionation (more specifically
coarsening) of the drop size distribution.

(i) LSW theory. Theory for the change in droplet size distribution owing to Ostwald
ripening in emulsions has been developed by Lifshits and Slezov^^ and independently

^^ See, e.g., S.E. Friberg, P. O. Jansson, and E. Cederberg, J. Colloid Interf. Sci. 55 (1976) 614.
For the most part, based on reviews by A. Kabalnov, E.D. Shchukin, Adv. Colloid Interface
Sci. 38 (1992) 69; J.G. Weers, chapter 9, p. 292, in B.P. Binks, Ed. (1998), General References;
and A. Kabalnov, J. Disp. Sci. Technol. 22 (2001) 1
^^ After I.M. Lifshits, V.V. Slezov, Zhm.Eksp. Tear. Fiz. 35 (1958) 474; transl. in Soviet Physics
JETP 35 (1959) 331. The names arc also transcribed as Llfshitz, Slyezov and Slyozov.
8.66 EMULSIONS

by Wagner ^^ This LSW theory proceeds on the following assumptions:


- the drops are fixed in space;
- the distance between drops is very much larger than d , which implies that ^ -> 0 ;
- the concentration of the material of the disperse in the continuous phase is
everywhere the same (corresponding to the solubility of the material in drops of radius
a = a^j., explained below), except near the drop surfaces where a local concentration
gradient exists.
Ostwald ripening then results after some time in a number frequency distribution
of the drop radius given by

u^exp[3/(2u-3)]
Jin) ~ (u+ 3)^/3(1.5-u)l 1/3' 0<u<1.5 [8.3.2]

where u = a/a^^; a^^ is the radius of a drop that, at the moment considered, neither
shrinks nor grows. The size distribution is of a fixed shape, shown in fig. 8.23b.
The change in a^^. is now given by.

d < _^^'-P^sato. [8.3.3]


dt 9PD

where x' is given by [8.3.11 and D is the diffusion coefficient of the drop material in
the continuous phase; c^^^is in kg m"*^. Equation [8.3.3] holds for any type of average
radius, though with a different numerical constant. The time needed to double the
average volume of the drops would be given by CL^^-Q divided by the r.h.s. of [8.3.3].
Figure 8.23a illustrates changes in droplet size. Small droplets disappear, and hence

/(u)

Figure 8.23. Ostwald ripening according to the LSW theory, (a) Examples of the change in
radius (a) with time {t) of some drops in an emulsion; the heavy line gives the change in a^^.
(b) The normalized distribution /(u) .

' C. Wagner, Z. Elektrochem. 35 (1961) 581.


EMULSIONS 8.67

the number of drops per unit volume, N^ , decreases. After the LSW size distribution
has been formed, the decrease is by a second order reaction, and is given at any time
by
x'Dc^
-Ni [8.3.4]
dt PD

with a numerical constant close to unity.


Table 8.3 gives some calculated examples of the rate of Ostwald ripening. The rate
is proportional to a^ and to D, and inversely proportional to y and to c^^^ ^. The
latter especially is an important variable. For example, emulsions of hexadecane or
triglyceride oil in water [i.e., oils that are virtually insoluble in water), show negligible
Ostwald ripening, despite the driving force being substantial.

Table 8.3. Calculated examples of the time t * needed to double a^^ due to Ostv^^ald
ripening according to LSW theory (uncorrected). Pe is a Peclet number; see [8.3.5].

Disperse phase Benzene Decane Tetradecane Water


Continuous phase Water Water Water Triglyceride oil |

«cr.o/^^n^ 5 0.5 0.5 1

1 r^^/mNm-i 35 10^^ 10"^ 10"^

Mj^/Da 78 142 198 18

p^/kgm-^ 880 730 763 998

x'/nm 2.5 1.6 1.9 0.15 1


D/m^s-i 10-9 8x10"^^ 8x10-10 2x10-11
Sat.oo/kgm-3 0.7 5x10-5 2.8 xlO-'^ 1.4

t* 39 hours 37 days 16 years 6 days

Pe 1.4 14 12 150

Presence of surfactant
The LSW theory is confirmed experimentally, insofar as it concerns the effects of
the variables in the equations; the linear increase of a^ with time is also well obeyed.
However, the absolute rate is generally substantially higher than that predicted, e.g. by
a factor of 5. It is difficult to establish quantitative relations experimentally. In order to
obtain results in a reasonable time, very small drops ( a down to 0.1 |Lim) are often
used, and it is then quite difficult to obtain reliable size distributions. IVloreover, /
may vary during the process.
The main factors identified as being responsible for deviations from LSW theory
are:
8.68 EMULSIONS

- The drops are not fixed in space, but show Brownian motion. Whether this affects
the ripening rate can be derived from the value of a translational Peclet number for
particle diffusion over molecular diffusion. It is given by^^

Pe^—(3kT/mf^^ [8.3.5]

where m is the droplet mass, and D the molecular diffusion coefficient. If Pe >\,
Brownian motion will enhance Ostwald ripening. Table 8.3 also gives value of Pe and,
except for the first example, its value is considerable. Quantitative relationships for the
ripening rate have not been given, but enhancement by a factor of at least 2 has been
observed experimentally. Convection in the emulsion, e.g. due to temperature
fluctuation, can further enhance the ripening rate.
- The assumption that the distance between drops is much larger than their
diameter only holds for very low volume fractions. A theoretical treatment^^ resulted in
an increase of ripening rate by factors of 1.4, 1.75, and 2.2 for (p = 0.03, 0.1 and 0.3,
respectively. In some experiments, this fitted the results fairly well, in other cases, less
so.
- If the surfactant present forms micelles, these may solubilize material from the
disperse phase, say, an oil. It is still a matter of debate whether transport of the
material will then be enhanced and, if so, by what mechanism. For non-ionic
surfactants, enhanced ripening by micelles has been well established, but not for
ionics. If an excess of surfactant is added to an emulsion, it can be observed that the
number of droplets at first decreases without a marked increase in average droplet
size; this is because the micelles take up a significant amount of oil (until they have
reached their maximum swelling).
- The LSW equations [8.3.3-4] only apply after the size distribution of [8.3.2] has
established itself, which may take considerable time. If the initial size distribution is
quite narrow, initial Ostwald ripening is slower than predicted. Most emulsions, how-
ever, have wide size distributions, the relative standard deviation being (much) larger
than that of f{u), which is about 0.25. In such a case, the initial ripening rate will be
greater.

(ii) Effects of the surfactant. The foremost effect is that a surfactant lowers y,
which will proportionally decrease the ripening rate. Some surfactants can cause a
very low interfacial tension (see sec. 8.1b, sub (i)), greatly slowing down Ostwald
ripening, but the droplets are then generally prone to rapid coalescence. If y^^ is
larger than its plateau value, Ostwald ripening will lead to a lowering of y, since the
decrease in specific interfacial area can result in a greater surface excess. Another

^^ A.S. Kabalnov, K.N. Makarov, A.V. Pcrtsov, and E.D. Shchukin, J. Colloid Interface Set 138
(1990)98.
^' Y. Enomoto, K. Kawasaki, and M. Tokuyama, Acta Metall. 35 (1987) 907.
EMULSIONS 8.69

effect may be that several non polymeric surfactants can form micelles, which possibly
enhances mass-transport between drops; see above.

Important effects can be caused by the action of a surface dilational modulus, K J .


Using the Laplace pressure as a variable, Gibbs derived that Ostwald ripening will
stop if

dpL ^ d(2y/a) _ 2 a ( d r / d a ) - 2 y _ 4Kg - 2 / ^ ^ ^^^^^^


da da a^ a^

where use has been made of the relationship din A = (2/a)da . The condition can be
written as,

K^>y/2 [8.3.6b]

This conclusion should hold if the following conditions are fulfilled:


- The larger drop is only slightly larger than the smaller one; in other words, the
drop size distribution is very narrow. In practice, the condition will be more like

- The value of KJ remains constant in time, implying that the surface is covered by
a true Langmuir monolayer (i.e., KJ is fully elastic). This is rarely the case, and the
consequences will be discussed below.
- The drop will remain spherical. However, the surfactant film of a shrinking
droplet is being compressed and it will probably buckle; the drop will then collapse
into a dented particle. The condition for buckling is roughly given by,

a>SK^/r [8.3.7]

where S is the thickness of the adsorption layer. Assuming this to be 5 nm, and the
value of K^/y to be 2-5, as is often the case, a droplet of radius over 10-25 nm will
tend to collapse. Droplet collapse has frequently been experimentally observed. For
some surfactants, the value of K j / y may become much higher on compression of the
monolayer, so that larger droplets can escape collapse. The result may then be
formation of a bimodal size distribution ^^
- The drops do not coalesce. The growth of drops covered by a Langmuir mono-
layer will result in a decrease of their surface excess F; if F remains low, it may
promote droplet coalescence, as has indeed been observed'^^
Altogether, stopping Ostwald ripening by application of a surfactant that gives a
high surface dilational modulus will rarely be successful. However, Ostwald ripening
can, in principle, be stopped by making droplets that are covered by suitable solid
particles, as applied in Pickering stabilization against coalescence. The particles will

^^ See, e.g., M.B.J. Meinders, W. Kloek, and T. van Vliet, Langmuir 17 (2001) 3923.
^^ E. Dickinson, C. Ritzoulis, Y. Yamamoto, and H. Logan, Coll. Surf. B12 (1999) 139.
8.70 EMULSIONS

not desorb upon shrinking of the droplet, and the fairly thick and stiff layer will not
readily buckle ^^

In nearly all cases the surface dilational modulus is a complex number (see [8.1.5]),
implying that the surface is viscoelastic. This may result in a substantial decrease in
Ost^vald ripening rate. One may use either a time dependent modulus^^ or a strain rate
dependent surface dilational viscosity^^ to calculate the ripening rate, but analytical
exprfvssions are not available. Numerical methods have yielded results that agree
rather well with experimental results. For emulsions stabilized with proteins, the rate
may be decreased by a factor up to five'*^ This applies especially to those globular
proteins that exhibit inter molecular cross linking after adsorption.
Bulk rheological properties may also affect the ripening rate^^ (apart from the effect
of continuous phase viscosity on the diffusion coefficient). A bulk yield stress larger
than the Laplace pressure of the smallest droplets will prevent droplets shrinking or
growing in size. However, the yield stress may have to be quite high, e.g., 1 MPa, which
is difficult to realize.

(Hi) Drops of mixed composition. If the disperse phase consists of two components
of different solubility in the continuous phase, this may retard Ostwald ripening,
owing to the development of a counteracting change in chemical potential. A simple
example is given when one of the components is not soluble in the continuous phase,
such as aqueous drops containing some salt in a triglyceride oil. Upon shrinking of a
drop, its molar salt concentration c (in osmolm"^ )^^ increases over the initial value
CQ , whereby the osmotic pressure in the drop increases according to

^osm = '^R^ = C o K / « f « ^ 18.3.8]

assuming ideal behaviour. No change in droplet size will occur if the increase in
Laplace pressure owing to shrinkage equals the increase in H^^^ . This point is
reached if

Y=l.5acRT [8.3.9]

Assuming 7 = 10mNm"^ and a = 1 |j.m, a value of c = 2.67 osmolm"^ would be

For foam bubbles this has been shown clearly by Z. Du, M.P. Bilbao Montoya, B.P. Binks, E.
Dickinson, R. Ettelaie, and B.S. Murray, Langmulr 19 (2003) 3106.
^^ M.B.J. Meinders, M.A. Bos, W.J. Lichtendonk, and T. van Vliet, in E. Dickinson, T. van Vliet,
Eds., Food Colloids, Biopolymers and Materials. Royal Soc. Chem. (2003).
^^ A.D. Rontcltap, B.R. Damste, M. dc Gee, and A. Prins, Coll. Surf. 47 (1990) 269.
See. E. Dickinson, et al., loc. cit.
^^ W. Kloek, T. van Vliet, and M. Meinders,
M( J. Colloid Interface Sci. 237 (2001) 158.
The osmolarity is the osmotically
osmotical active concentration; for instance, for ideally dilute NaCl
and MgCl2 solutions it equals 2C^^QY ^^^ 3cjyjgQj , respectively.
EMULSIONS 8.71

sufficient to stop ripening; this corresponds to less than 0.01% NaCl. The presence of
a 'trapped component', i.e., a component soluble in the disperse phase, but not in the
continuous phase, can thus be an effective means of stopping Ostwald ripening.
Equations for the rate of Ostwald ripening can be derived if all drops contain two
components, 1 and 2, which both are soluble in the continuous phase. Assume that
the volume fractions of the components in the original drops are /^ and / 2 f .- lere
/ j + / 2 = l ) . When the condition /2<C2^/Cj^ is fulfilled, the rate would be,
according to LSW theory, given by

dt 9p2

The rate is thus given by the scale x' for component 1, and the solubility and
diffusion coefficient of component 2. If the mentioned condition does not hold, the rate
is given roughly by [8.3.3], calculated for component 1.

If an emulsion consists of two kinds of drops containing different materials, which


are both soluble in the continuous phase, then compositional ripening will occur ^^
(besides Ostwald ripening). The materials are exchanged between drops and, even-
tually, all drops will attain the same mixed composition.

8.3c Aggregation
Colloidal interaction forces between particles, and the kinetics of aggregation, have
been discussed extensively in chapters IV. 3 and 4, and additional information is in
chapters V. 1, 2 and 3. Hence, we need not discuss aggregation in detail. In a sense,
emulsions are ideal systems to study aggregation, although the polydispersity and the
difficulty of estimating the colloidal interaction forces (see sec. 8.1g) may pose some
problems. We will only mention points that are specific for (some types of) emulsions.
Kinetics. Large drops, say > 3 |im, do not generally show perikinetic aggregation.
Small velocity gradients resulting from temperature fluctuations (see [8.3.17]) can
strongly promote orthokinetic aggregation. Assuming that the capture efficiency is the
same in both cases, the ratio of the rates of orthokinetic over perikinetic aggregation is
given by,

where Vv is the velocity gradient. In water at room temperature, drops of d = 3 [iim


will aggregate faster by a factor of three, owing to a Vu of only 1 s~^. IVIorcover, the
drops will generally show sedimentation and, owing to their polydispersity, the larger

^^ See, e.g., L. Taisne, P. Walstra, B. Cabane, J. Colloid Interface ScL 184 (1996) 378.
8.72 EMULSIONS

drops overtake the smaller ones; this will strongly enhcince the encounter frequency.
Film Jormation. When two emulsion drops form an aggregate and the attractive
forces are relatively strong, a flat film may form between the drops. This means that
the net force keeping the drops together increases. This phenomenon will not, or will
hardly affect the aggregation rate, but it may well decrease the possibility for deaggreg-
ation, for example, owing to a velocity gradient in the liquid. Film formation upon the
encounter of drops will be discussed in sec. 8.3e.
Dielectric permittivity of the continuous phase. The dielectric constant in the oil of
a water-in-oil emulsion can be quite small, say 3. The water drops may have a surface
charge, but the decay of the electrostatic potential with the distance from the drop
surface is very weak; in other words, l//r is very large. Consequently, the repulsion is
almost purely Coulombic. Since the variation with distance of the repulsive force is
quite low, it provides very little stability to the drops. These aspects were discussed in
sec. IV.3.11. It is often difficult to find small molecule surfactants that stabilize w/o
emulsions; suitable copolymers will provide steric repulsion.
Depletion interaction (see sees. 1.8 and 9). Substances considered essential for
making stable emulsions may well cause depletion interaction, and hence droplet
aggregation. To slow down sedimentation, high polymers are often added to increase
the continuous phase's viscosity, but if they are non-adsorbing, they may also cause
depletion interaction. For example, the addition of 0.02% xanthan gum may already
cause aggregation, hence rapid sedimentation. Higher polymer concentrations may
lead to formation of a particle gel; aggregation is then so fast that a space filling
network of drops is formed before appreciable sedimentation occurs.
Depletion interaction can also be caused at a high amphiphile concentration, when
the amphiphiles form micelles ^^ The depletion Gibbs energy per pair of drops is given
roughly by

AGdepi = - 2 7 i a d / 7 ^ ( 2 a „ - h f . 0<h<2a^, a<j » a^


18.3.12]

where a = radius, h = distance between drops, 77 = osmotic pressure, c = concen-


tration (mass/volume), M= molar mass, and (p= volume fraction; the subscripts d
and m refer to drops and micelles, respectively. It is seen that the drop radius and the
micelle concentration, especially, have a large effect on AG. The interaction is
strongest for h = 0 , but this value may not be attained if there is strong repulsion
between drops at very small h . To give an example, a concentration of 1% of sur-
factant in the aqueous phase will readily cause depletion aggregation of 2 |im drops
(AG == -4kT]. For a more elaborate discussion, see sees. 1.8 and 9.

See, e.g., M.P. Aronson, in J. Sjoblom, Ed., Emulsions: A Fundamental and Practical
Approach, Kluwer Academic (1992), p. 75.
EMULSIONS 8.73

Mixtures of non-ionic amphiphiles and caseinate have also been incriminated as


causing depletion flocculation in an emulsion ^^
Bridging. Small particles, or long polymer molecules, that adsorb onto the drops,
but are preferentially wetted by the continuous phase, can be adsorbed simultaneously
onto two drops, thereby forming a bridge between them. An example is given by bridg-
ing by protein aggregates in o-w emulsions. This can occur especially if the mentioned
species form the sole surfactant and are present in insufficient quantity to fully cover
the droplet surfaces. Bridge formation can then generally occur during emulsification,
but rarely afterwards. The addition of some suitable amphiphile to the emulsion can
cause desorption of the particles, and stirring then leads to deaggregation.
Other bridges can form by reactions between specific groups of the adsorbed sur-
factants, especially with polymers. After all, steric repulsion caused by polymers does
not preclude temporary partial overlap of the adsorbed layers. An example is found in
the bridging by divalent cations associating with monovalent acid groups, or covalent
-S-S- cross links between adsorbed proteins. These bridges generally form after
emulsification.
Partial coalescence. Crystals can form in some emulsion drops, for example, if the
drops consist of triglyceride or paraffin oil. These oils are generally multicomponent
mixtures and the proportion crystalline will greatly depend on the temperature and its
history. Often, some crystads protrude a little from the droplet surface. Such a crystal
can pierce the film between approaching droplets, and thereby form a bridge. If the
contact angle is suitable, say between 110 and 170° as measured in the aqueous
phase, the crystal will be wetted by oil. If so, there is oil oil contact between drops, and
a driving force for complete coalescence. However, the latter will generally be prevent-
ed, since the crystals in a drop tend to form a space-filling network. Hence the term,
'partial coalescence' (also called 'clumping'). Heating the emulsion to cause the crystals
to melt leads to the coailescence of each clump into a large drop.
In a quiescent emulsion, aggregation due to partial coalescence is generally very
slow, but on application of a velocity gradient the process can be very fast, millions of
times faster than true coalescence of a similar emulsion without crystals. In simple-
shear flow, the drops will roll over each other upon encounter, greatly enhancing the
chance of a crystal piercing the film; piercing may also be promoted by the shear
forces pressing the drops closer to each other. The rate and the consequences of
partial coalescence depend in an intricate manner on several variables^ .

8.3d Sedimentation^^
By the buoyancy principle of Archimedes, the net gravitational force acting on a
single sphere of diameter d submerged in a liquid is given by

^^ E. Dickinson, C. Rltzoulis, and M.J.W. Povey, J. Colloid Interface Set 212 (1999) 466.
^^ For a review see P. Walstra (1996).
For the most part based on P. Walstra (1996), General References.
8.74 EMULSIONS

Fi,=-^Kd3g(^^-pJ 18.3,131

wher< g is the acceleration due to gravity. Hence, the sphere will move downward
thro igh the continuous liquid if the density difference is positive, or upward if it is
negative. The moving sphere will be subject to a frictional drag force, according to
Stokes, given by

Fj =Jv = 3nd7]^v [8.3.14]

where / is the friction coefficient and v is the linear velocity of the sphere with
respect to the surrounding liquid. Soon after the drop starts moving, inertial forces
will become negligible and the drop will sediment at a constant 'Stokes velocity'. This
is obtained by putting F^ = F^:

» s = ^ 18.3.151

For example, an oil droplet of 2 |im in water, Ap = -70 kg m"*^, will attain a sedi-
mentation rate v^ = - 0 . 1 5 )am s~^, at room temperature, i.e., cream by 13 mm per
day.
Equation [8.3.15] can be modified for centrifugal sedimentation by replacing g by
the centrifugal acceleration, Rcoi^ , where R is the effective centrifuge radius and co
the rate of revolution in rad s~^.
The Stokes equation is frequently used for calculating sedimentation rates, but is
only valid under a very restricted range of conditions.

(i) Prerequisites for the Stokes equation. The following conditions should be met
for the equation to hold:
- It concerns one sphere in an infinite amount of a homogeneous liquid. In practice
this means that the volume fraction of the emulsion drops must be very small, say
^ < 0.003, and the vessel should have a diameter at least 50 times d. In most
emulsions, hindered sedimentation occurs, which is discussed below.
- The drops must be homogeneous perfect spheres. IVIany emulsion drops nearly
meet this condition, but if the drops are quite small and covered with a polymeric
surfactant, complications may arise. It is often possible to calculate approximate
corrections.
- The drop surface must be immobile in the tangential direction. As discussed in
sec. 8.1c, this is nearly always the case.
-The drop Reynolds number, given by dvp^/j]^, must be much smaller than
unity. This is virtually always the case, also for centrifugal sedimentation.
- Brownian motion should not disturb the sedimentation. For very small drops, the
EMULSIONS 8.75

Stokes equation is no longer valid and a sedimentation equilibrium is (slowly) estab-


lished. The concentration, c , of the droplets as a function of the vertical distance, z ,
is then given by

c,z, = C o e x p [ z ! i | g M j ,8.3.16,

Assuming the disturbance owing to Brownian motion to be negligible if


C{Z)/CQ < 1.02 , and taking Ap = - 7 0 kg m~^ and the maximum z value at 0.2 m, the
condition becomes d > 50 nm , which will nearly always be fulfilled.
- Convection currents should not disturb sedimentation. Such currents can form,
due to slight temperature fluctuations. To illustrate this, the free convection flow
between two vertical plates with temperature difference, AT, at a mutual distance, x,
will be considered. The average velocity gradient is given by^^

Vu = ^ ^ ^ ^ ^ ^ 18.3.171
32 77^
where /3 is the volume expansion coefficient. For water at room temperature and
X = 0.05 m , Vi; would roughly equal SAT; a temperature difference of 0.1 K would
then produce a gradient of 0.3 s~^. The stress caused by a gradient will equal
VLJ X T;^ , and it should be much smaller than the net gravitational stress, which equals
4F^ /Tcd^. This leads to the condition,

677^
d»Wv ^ [8.3.18]
ng\A/o\

implying that for oil drops in water, the value of d , in |im, should be much larger
than 2 Vu in s~^. For a AT value of 0.1 K, this implies d » 0.6 |im . This is a more
restrictive criterion than the one for negligible Brownian motion. It is indeed observed
that sedimentation of 1 jim droplets in water hardly occurs, unless the temperature is
kept precisely constant.
- The continuous phase should be a Newton liquid, implying that the viscosity is
independent of the viscous stress (in a stress controlled rheometer) or of Vi; (in a
shear rate controlled instrument). In many emulsions, the liquid is not Newtonian, the
apparent viscosity T]^ generally decreasing with increasing stress. The stress acting on
a sedimenting drop will, to a first approximation, be equal to F^ divided by the
droplet surface area, i.e., dgf|Ap|/6 . For an oil drop of 5 jim size in water, this stress
would then be about 6x10""^ P a . The viscosity used to calculate the Stokes velocity
should therefore be measured at that value of the shear stress. However, most
rheometers cannot give results at a stress below 0.1 or even 1 Pa, and the apparent
viscosity measured may then be a considerable underestimation of the value prevailing

^^ R.B. Bird, W.E. Stewart, and E.W. Lightfoot, Transport Phenomena, Wiley (1981) p. 217.
8.76 EMULSIONS

during sedimentation, by up to some orders of magnitude. It should be noted that the


reasoning given here is an oversimplification, because the liquid flow around a
sedimenting droplet has an elongational component; especially for vlsco elastic liquids
this can make a substantial difference.
It may even be that the liquid has a yield stress that is greater than the net
gravitational stress acting on the drop, given approximately by dg|Ap|, but far smaller
than the smallest stress to be applied in the rheometer. Such a yield stress is not then
detectable, although it can completely prevent sedimentation.

(ii) Hindered sedimentation. The Stokes equation gives the velocity of the drop
relative to the surrounding liquid. Since we are generally interested in the velocity
relative to the vessel containing the emulsion, a correction for the displacement of
continuous phase must be made: the Stokes velocity Dg has to be multiplied by
(1 - ^ ) . However, much larger corrections are needed.
Batchelor attempted to derive a rigorous equation for homodisperse spheres of
q)«OA ^\ The result is

— = l-6.55(p 18.3.19]

The factor -6.55 is composed of the following three terms, (i) An amount -5.50 for
counterflow of continuous phase, owing to the drops' displacing liquid and, more im-
portantly, dragging liquid along with them; (ii) an amount -1.55 owing to the overlap
of the flow disturbances caused by drops sedimenting near to each other; (iii) an
amount +0.50 caused by group sedimentation (see below).
The relationship only fits (approximately) for very small cp values, and exponential
equations have been proposed for higher volume fractions, e.g.^^

— ^{l-tpf 18.3.20]

where the exponent n should presumably equal 6.55. This equation fits for far larger
(p values, but the exponent tends to be significantly larger: see, e.g., fig. 8.24, solid
line, where n = 8.6 . Pubhshed values for n vary mostly between 7 and 9.
A major factor explaining the high values of the exponent is polydispersity. In a
theoretical study on mixtures of spheres of two sizes, again at low (p'^\ it was
concluded that, as expected, the value of v/v^ increases for the smaller spheres, and
decreases for the larger ones; however, the total sedimentation rate was calculated to
decrease. Experimental studies have also shown a decrease in sedimentation rate with
increasing polydispersity.

^' G.K. Batchelor, J. Fluid Mech. 52 (1972) 245.


2' J.F. Richardson, W.N. Zaki, Trans. Inst. Chem. Eng. 32 (1954) 35.
^^ G.K. Bachelor, C.S. Wen, J. Fluid Mech. 124 (1982) 495.
EMULSIONS 8.77

log (u/u. Figure 8.24. Creaming rate relative


to the Stokes velocity v/v^ of mix-
tures of skim milk and cream, of
various oil volume fractions g?.
Average droplet size, d^^ = 1.5 ^im .
( • ) Creaming under gravity; (o) ibid,
in a centrifuge at 200 xg .

-0.4
0.02 0.04
-log(l-(p)

The average drop size that should be taken for polydisperse systems is d^g , as is
explained in sec. 8.1e.

(Hi) Group sedimentation. Figure 8.24 gives creaming results on a series of o/w
emulsions of various (p, both under gravity and in a centrifuge. The conditions
(geometry, and time of creaming) were chosen so that both sets of experiments would
give the same total creaming according to the Stokes equation. All of the prerequisites
for the Stokes equation mentioned in subsection (1), except the first one, were fulfilled,
so that It would be expected that values extrapolated to ^ = 0 would be equal for the
two sets. This was indeed observed. It Is also seen that [8.3.20] Is well obeyed for
gravity creaming, but that the results for centrifugal creaming were very different. This
must be related to differences in group sedimentation.
As was mentioned In subsection (11), Batchelor has considered this phenomenon. If
two drops are quite close to each other, the total friction factor of the doublet was
calculated to be smaller than twice that of a single drop, especially if the one drop is
above the other. If so, the sedimentation rate will be enhanced. Batchelor assumed
that the mutual positions of the drops would not change during the process, but that
is unlikely. Drops diffuse, i.e., they move by Brownian motion, which may bring them
into a mutual position in which sedimentation is enhanced. In principle, this would
increase the overall sedimentation rate. Whether the increase is significant depends on
the time scales involved, i.e., whether the time during which two drops are close to
each other is long enough for a substantial increase in sedimentation to occur. It
should be taken into account that the sedimentation distance is proportional to time,
whereas the distance of diffusion is proportional to the square root of time. The
tendency for group sedimentation to be significant may be expressed in a translational
Peclet number, i.e. the ratio of the time needed for a droplet to diffuse over a distance
equal to d, over that for sedimentation over the same distance. From the Einstein and

^^ After results by P. Walstra, H. Oortwijn, Neth. Milk Dairy J. 29 (1975) 263.


8.78 EMULSIONS

Stokes equations we derive,

Pe = ^J-^ [8.3.21]
6kT
for gravity creaming; in a centrifuge, g should be replaced by Ro}^ . If Pe » l , group
sedimentation will be of importance. In the example of fig. 8.24, Pe ~ 0.5 for gravity
creaming and about 100 in the centrifuge. It is seen in the figure that near (p = 0.03
the 'centrifuge' curve shows a maximum, where v is about 5/3 the value for gravity
creaming. If gravity creaming concerns larger drops, say some micrometers in
diameter, Pe will also be large. For example, in dispersions of small glass spheres in
water, an increase over the Stokes rate by a factor of 2.3 was observed^^ near
^ = 0.03.
Precise prediction of the sedimentation rate at values of (p that are not very small
is not easily possible. The degree of polydispersity may also affect group sediment-
ation. It is fairly obvious that group sedimentation will be of little importance at very
low (p, since transient doublet formation will be rare. The higher is (p, the stronger is
the effect, but the mechanisms that cause a decrease in sedimentation rate will also
have a greater effect. At very high volume fractions, the notion of group sedimentation
becomes irrelevant.

(iv) Colloidal interactions. So far, it has been implicitly assumed that colloidal
interaction forces between the drops are negligible (except for hard core repulsion). Of
course, such interaction forces exist, but the range is often quite small compared to
the drop radius. Nevertheless, for cases where the range is relatively long (small drop-
lets, electrostatic repulsion at low ionic strength), net repulsive interaction decreases
the sedimentation rate somewhat^\ presumably owing to decreased group sedimenta-
tion. For weak attractive interactions, one would expect an enhanced sedimentation
rate.
For stronger attractive forces, irreversible aggregation will occur, and this can
greatly affect the sedimentation rate. If the drops form fractal aggregates, as will often
be the case, the sedimentation rate of an aggregate relative to that of a single drop will
be given by^^

'^aggr=Vopl-8N("/<») 18.3.221

where N i s the number of (equal sized) drops in the aggregate, and d is the fractal
dimensionality (which will generally be about 2); the proportionality^ factor 1.8 is
approximate, and depends somewhat on the value of d. To give an example, for
iV = 100 , i; would be increased by a factor of 18.

^' R. Johne, Chem. Ing. Techn. 38 (1966) 428.


^' E. Dickinson, J. Colloid Interface Set 73 (1980) 578.
^' L.G.B. Bremer, Ph.D. Thesis, Wageningen University, 1992.
EMULSIONS 8.79

In practice, the s i t u a t i o n is m o r e complicated. T h e d r o p s are p o l y d i s p e r s e . Larger


d r o p s will overtake s m a l l e r o n e s , leading to e n h a n c e d aggregation. Larger aggregates
will o v e r t a k e s m a l l e r o n e s a n d single d r o p s , which will e n h a n c e aggregation even
m o r e . Altogether, the r a t e of aggregation a n d s e d i m e n t a t i o n will i n c r e a s e ever faster,
a n d a s e d i m e n t or c r e a m layer is rapidly formed. See also fig. 8 . 2 5 .

0.8

w
^^^•^s^
(a) (b)
60
0.4 h L ^^ 6 2

0
2 6
y<::z^^^^—^J

0 1
d i s t a n c e from b o t t o m t/%0

Figure 8.25. Examples of the concentration profile developing during creaming of emulsions.
The drop concentration is given as volume fraction, ^ , as a function of height, z. The numbers
near the curves denote the time after creaming started (say, in hours), (a) Calculated for a
strictly homodisperse emulsion, (b) Polydisperse emulsion, no aggregation, (c) Polydisperse
emulsion with aggregating drops, (d) Percentage of the total drop volume creamed as a function
of time t, for cases a c; t^Q is the time needed for half of the total drop volume to reach the
cream layer. Approximate examples, meant to illustrate trends .

If t h e d r o p s a r e n o t too large, v is not very small, a n d aggregation is relatively fast,


fractal aggregation will lead to the formation of a s p a c e filling particle n e t w o r k , i.e., a
gel. T h i s will s t o p s e d i m e n t a t i o n . However, gravity m a y c a u s e c o n s o l i d a t i o n of the gel,
a n d t h e n a layer of c o n t i n u o u s p h a s e will b e formed at the t o p or t h e b o t t o m of the
vessel.

fuj Sedimentation profiles. In practice, one is often i n t e r e s t e d in the d e v e l o p m e n t


of a s e d i m e n t a t i o n profile, a s k i n g w h a t is the volume fraction of d r o p s a s a function of
height in the vessel a s a function of time? T h i s will be illustrated for creaming, with
reference to fig. 8 . 2 5 .
If all d r o p s are of t h e s a m e size, a s h a r p d e m a r c a t i o n plane f o r m s b e t w e e n a liquid

From P. Walstra, 2003, Physical Chemistry of Foods, Marcel Dekker (2003).


8.80 EMULSIONS

with (p = 0 and one with (p= (pQ (the original value in the emulsion). The plane moves
upward; if (pQ<0.\, at a velocity according to [8.3.20]. A cream layer forms with
(p~ 0.7, and the plane between emulsion and cream moves downwards. During
creaming, the value of i; between the two planes remains constant. This is illustrated
in fig. 8.25a.
The situation is far more complicated for a polydisperse emulsion. The volume
fraction of drops reaching the cream layer per unit time will be given by

q^ ,,^^' '/M [8.3.23]

where z^^ is the creaming distance and J[(p) includes all the corrections to the Stokes
velocity as a function of the local value of cp. The main problems are that the values of
(p, dgg , and polydispersity vary with time and z ; the creaming distance decreases as
the cream layer becomes thicker. Figure 8.25b illustrates an experimental result. Such
results can be obtained by scanning the creaming vessel, as a function of z , with an
ultrasound beam and a detector; the ultrasound attenuation can be related to the local
volume fraction ^\
Figure 8.25c illustrates what can happen if the droplets aggregate during creaming.
Now, a fairly sharp demarcation plane can form between the emulsion and continuous
phase,, and the creaming rate increases with time, contrary to the situation where no
aggregation occurs, when creaming becomes ever slower. This is further illustrated in
fig. 8.25d.
The figure also shows that the cream layer can become compacted after its initial
formation, especially in the case of aggregating drops.

8.3e Coalescence^^
Coalescence involves some successive phenomena: two drops encounter each other;
a film is formed between them; the film has become thinner; and if the film becomes
quite thin it can rupture. The last named event immediately induces coalescence of the
two drops.
Several theories have been developed to predict the time it takes, on average, for a
single film of given properties to rupture'^^ Rather than the time, its reciprocal, the
(hypothetical) film rupture frequency, J^ (unit s~^) is given, often in the form,

Jf = ^ a ^ exp(-AG^^^ /kT) [8.3.24]

^ See, e.g., M.J.W. Povey, in E. Dickinson, Ed., New Physico chemical Techniques for the
Characterization of Complex Food Systems, Blackie Academic (1995), chapter 8, p. 196.
^^ Largely based on Walstra, 1996; on chapters 1 (by B.P. Binks), 7 (by A.S. Kabalnov), 8 (by B.
Dcminiere, A. Colin, F.L. Caldera, and J. Bibctte) and 10 (by D.N. Petsev) in Binks (1998); and
on LB. Ivanov, P.A. Kralchevsky (1997); see General References.
For the thinning and rupture of isolated planparallel liquid films, see sec. 6.4c.
EMULSIONS 8.81

where A^ is the area of the film and AG^^^^ is an activation Gibbs energy for rupture.
The natural frequency v^ (units m"^ s^M is the presumed rupture rate if AG = 0 . It
is often supposed that it equals the frequency at which a single surfactant molecule
leaves the adsorption layer. Applying the absolute rate theory for molecular reactions,
the frequency for an area occupied by a single molecule would be given by ,

v^ =kT/hz'^ [8.3.24a]

where z is the molecular dimension. Another theory^^ is based on the fact that
transverse thermal waves develop spontaneously on a liquid interface (as follows from
surface light scattering studies; sec. III. 1.10), and assumes that a wave of length equal
to z may cause the removal of molecules from the surfactant layer. The frequency co
then would be in the ultrasound region, and Q) = V^/Z where D^ is the sound velocity.
The occurrence per unit area of one molecule, z^ , is then given by

v^ = (0/z^ =v^/z^ [8.3.24b]

Assuming that u^ = 1 km s'^ and z = 1 n m , [8.3.24a] yields v^ - 10^^, and [8.3.24b]


about 10^^ m~^ s~^. It is not clear to what extent the equations are correct, since it is
difficult to obtain reliable data for AG^^,^. Nevertheless, [8.3.24] gives the main factors
determining the rupture frequency, and especially that it will be proportional to the
area of the film.
In practice, one is often interested in an overall coalescence rate, expressed, for
example, as the decrease in the number of drops per unit volume, N, with time. Plots
of 1/JV against time are often reasonably linear. Another method involves the
determination of the average size of the drops, e.g. d^2» ^^ ^ function of time.
Numerous factors affect the coalescence rate, and it is not possible to characterize
the overall situation v^th a set of simple equations. This is mainly because several
regimes for coalescence can be distinguished, and the main variables determining
coalescence rates are not all the same in the various regimes.

(i) Regimes. Rather than defining a great number of regimes, five variables are
given whose values are used to distinguish regimes.
a. Time available for film, rupture. When the drops are aggregated, or closely
packed as in a sediment layer, the time available is as long as the emulsion is kept.
Short times prevail when the drops remain separate until they encounter each other
by Brownian motion (perikinetic), or as a result of agitation (orthokinetic). The time
span during which two drops are then close to each other depends on several factors,
such as droplet diameter, colloidal interaction forces, and hydrodynamic conditions;
but it is generally below 1 s. For small droplets (order of a jim ) and in the absence of

^^ Due to P.G. de Gennes, J. Frost, The Physics of Liquid Crystals, Clarendon Press (1993)
p. 597.
8.82 EMULSIONS

colloidal Interactions, Brownian motion would, on average, cause the droplets to be


within a distance of roughly 10 nm for a time of the order of (in SI units) 10^ times
Tj^d . For droplets of 1 ^im in water, this would amount to 10 ms.
In practice, most emulsions with drops below 10 |im in diameter, which have been
made with a suitable surfactant and do not show aggregation, will not exhibit
coalescence, provided that sedimentation is prevented; the latter can be effected by
slowly rotating the vessel during storage.

Table 8.4. Role of the Weber number ([8.3.25]) in the deformation of emulsion drops;
calculated examples.

Laplace pressure
/ = 12mN m-l a = 0.25 |im PL-lO^Pa
y= 3 mN m~^ a = 6 fim PL - 10^ Pa

Local stress (7 = external stress x ia/h)

1. Van der Waals attraction^^ (T=A/I2nh^


A = 5x10-21 J, h = 10 nm o-==102pa
A = 5x10-2 J, h = 3 nm CT-5xlO^ Pa

2. Hydrodynamic shear stress (7= rjVva/h


77Vi; = 10Pa, a / h = 100 o-^lO^ Pa

3. Stress in a cream layer ^ ^ = Scream Vfif H a / h


Ap = 7 0 k g m - ^ , H = 1 0 m m , a/h = 100 ( 7 - 5 x 1 0 ^ Pa
Same in centrifuge at 1000 x g ( J - 5 x 1 0 ^ Pa

^^ It is assumed that strong repulsion occurs below distance h. A= Hamaker constant.


(p = volume fraction; Ap = density difference; H = height in cream layer.

b. Droplet deformation. Drops which encounter each other closely as a result of an


external force may become deformed, as illustrated in fig. 8.18. External forces may
be colloidal (net attraction), hydrodynamic, or gravitational. They are counteracted by
the Laplace pressure of the drops. The balance can be expressed in a dimensionless
Weber number, given by

We = ^^ = L_jexL = exl [8.3.25]


PL PL 2/h
where yl is a stress concentration factor that applies to undeformed drops, and which
equals the drop radius over the smallest separation distance, h; <j^^^ is the external
stress. For We «1, the droplets will not be deformed, meaning that the film area
remains quite small, of the order of a h . With increasing We, deformation gradually
becomes stronger and at We = l, a flat [i.e., parallel sided) film starts to form. For
EMULSIONS 8.83

We » 1 , a relatively large flat film is formed. Generally, coalescence occurs much


faster at We > 1 than at We « 1, other things being equal.
Some calculated examples of We values are given in table 8.4. It is seen that for
small drops and not very low interfacial tension, a« p^, hence We « 1, unless quite
a high externsil stress is applied.
c. Intensity of agitation. With an increasing agitation intensity of an emulsion, the
droplet encounter rate is enhanced (see [8.3.11]); the external stress acting on a
droplet pair increases (table 8.4, point 2); and the time needed for a film between
drops to thin to a given thickness increases for We » 1 , and decreases for We < 1.
Coalescence during intense agitation was discussed in sec. 8.2c, sub (vi); the con-
clusions obtained there also hold if the agitation is not so intense as to cause droplet
break-up. Altogether, the coalescence rate can be higher during agitation than in a
quiescent emulsion, £ind the relation between coalescence rate and some variables are
different (see below).
d. Interface rigidity. This is governed by the (transient) value of \K^\ , as discussed
in sec. 8.1c. The rigidity affects the rate of drainage of the film between drops, which is
determined by the Mcirangoni number {Ma, see [8.1.4]). From the discussion in sees.
8.1c and 8.3c, sub (vi), it can be concluded that in nearly all real situations Ma > 1,
implying that the interface is rigid. In extreme cases, e.g., an oAv emulsion with a very
low concentration of SDS and a low ionic strength (electrostatic repulsion will then
nevertheless prevent droplet aggregation), Ma can be small. Another, rather hypo-
thetical, case of a low Kg is the use of a very pure small molecule surfactant at a
very high concentration.
e. Surfactant category. There is a vast literature on coalescence theory, partly
backed up by experimental results, but in nearly adl of these cases it is (often im-
plicitly) assumed that the surfactant is a (small-molecule) amphiphile. Macromolec-
ular surfactants do not fit the theories; for example, [8.3.24a-b] for the natural
frequency of film rupture cannot apply, and the value of the film thickness, which
appeairs in aill film rupture theories, is unclear. For an emulsion stabilized with solid
particles the factors governing stability are different again.
All combinations of the two or three differences mentioned under each heading
would give rise to a very large number of regimes. Although several of these combin-
ations cannot, or hardly, occur, there still remains a substantial number of regimes to
be distinguished in practice.
It is often tried in practice to predict the coalescence stability of an emulsion from
the result of a test in which the coalescence rate is greatly enhanced. The prime
example is centrifuging the emulsion, or pressing the droplets together by osmotic de-
swelling of the emulsion. It may now be clear that this will generally alter the
coalescence regime by the two criteria (a) and (b). The change in Weber number,
especially, will cause the prediction to be unreliable.
8.84 EMULSIONS

(ii) Film formation and drainage. All theories for film rupture state that, for a
given system, the probability of rupture will be greater for a thinner film. Moreover,
the probability would be proportional to the film's area {[8.3.24]). Hence, film form-
ation and thinning are important aspects. For the thinning of isolated plane-parallel
films, see sec. 6.4a.
Film geometry. Whether a flat film is formed or not, is determined by the value of
the Weber number [8.3.25], but the interaction forces change when deformation
occurs. The complete equilibrium geometry of the film can in principle be calculated if
all interaction forces are known^^ Factors to be considered include; (i) the colloidal
interactions, e.g., van der Waals attraction, depletion interaction, electrostatic and
steric repulsion; and (ii) the deformation Gibbs energy, which is due to an increase of
interfacial area and to the bending energy of the surfactant monolayers (at the contour
lines of the flat film). For the latter, only the mean bending modulus need be taken
into account (see sec. III. 1.15), the Gauss curvature being negligible. The Gibbs energy
change owing to compression/dilation of the monolayers tends to be negligible. The
effect of a constant external stress can also be taken into account. The deformation
Gibbs energy is an important factor in determining the radius of the film formed, but
it does not affect the equilibrium thickness. The latter follows from the balance of the
colloidal and external interaction forces. If no repulsive interaction forces would act
(except hard core repulsion) a film of (nearly) zero thickness would be formed. Such a
'black' film can, in principle, be quite stable, but in most practical situations the film
would rupture before a black film has formed. It may be noted further that the
formation of a flat film implies the droplets' being irreversibly a^regated (as long as
conditions remain constant).
Drainage rate^\ The rate of drainage is discussed in sec. 8.2c, sub (vi). Equation
[8.2.21] would hold for We > 1, Ma > 1, and h«R«a (see fig. 8.18), a regime that
often prevails when coalescence occurs in an emulsion. In deriving the equation, one
should take into account that the force, F , and a are related, because the excess
pressure in the flat film between drops will equal the Laplace pressure of the drop;
s i n c e i R « a , p ^ will be close to that of the undeformed drop. Note, also, that the
drainage rate will be smaller for a larger F acting on the droplet pair. This is because
a larger value of F will cause a larger film radius, R , and hence more resistance to
drainage, especially for small values of h where drainage is hindered most. If the force
acting on the droplet pair is due to colloidal interactions, its magnitude will change
(considerably) with decreasing h , depending on the shape of the G{h) curve, and the
calculation of the drainage time is intricate.
For droplets in a simple-shear field, F will be, to a first approximation, propor-
tional to ri^^v, and hence constant. Integration of [8.2.21] for this case leads to an

A full account is given by D.N. Petsev, chapter 10 in Binks (1998), General References. See
also I.B. Ivanov, K.D. Danov, and P.A. Kralchevsky, Colloids Surfaces A152 (1999) 161.
^^ For the most part, after A.K. Chesters, Trans. Inst. Chem. Eng. 69-A (1991) 259.
EMULSIONS 8.85

expression for the time needed for the film to thin to a given value of h :

t{h) = • [8.3.26]
16/2h2

One should note the enormous effect of drop radius. (If the two drops are of unequal
radii, the relation 2/a = \/a^+\/a2 can be used.) Assuming that 77^=10"^ P a s ,
Vi; = 10^ s~^, and y = 3 mN m~^, t[h~5 nm) for spheres of a = 1 and 10 |im ,
would be 0.8 lis and 8 ms, respectively. The first conclusion may be that the times
are quite short. However, by comparing them to the time during which the drops are
close together (which is of order I/VLJ , hence 1 ms in the present case), it can be seen
that small drops readily come close to each other within that time, but large ones do
not. It is indeed generally observed that, for dilute emulsions that are agitated, the
coalescence rate is much slower for larger drops and a higher velocity gradient.
As discussed in sec. 6.4, large films [R> 50 |im ) often show uneven drainage:
sorne thin 'channels' and thicker 'islands' are formed. In such a case, [8.3.26] does not
apply.
Finally, the drainage rate is not always a relevant variable. For example, for small
droplets that form aggregates owing to perikinetic flocculation, or for drops that are in
a sediment layer, where the time available for drainage is quite long.

disperse phase

continuous phase
i- disperse phase

Figure 8.26. Cross-section through part of a film of (average) thickness h between two emulsion
droplets, (a) Illustration of hole formation, (b) Properties of a peristaltic wave developing in the
film.

(Hi) Film rupture. This subject is extensively discussed in sees. 6.4-6. Here, a few
aspects will be mentioned that are of special importance for coalescence in emulsions.
One characteristic of the film, then, is that it is quite small, the radius rarely being
larger than a few jim , and often as small as 0.1 |im .
De Vries theory. The first attempt to calculate the Gibbs energy for the formation of
a small hole or pore in a film was by de Vries ^\ Suppose that a hole has formed, as

' A.J. de Vries, Rec. Trav. Chim. 77 (1958) 383.


8.86 EMULSIONS

illustrated in fig. 8.26a. At point (1) in the figure, the Laplace pressure equals
Y(2/h-l/R); the two main curvatures at this point have different signs. At point (2),
Pj^ = 0 , and if the pressure at (1) is > 0, liquid will flow from (1) to (2) and the hole
will grow spontaneously. This can happen if 2/h>l/ R . To achieve this, the area of
the o/w interface has to be increased by an amount of the order h^ ; hence, the Gibbs
energy of the system would be increased by an amount yh'^ . The latter value can be
interpreted as the activation energy for hole formation, and hence for coalescence to
occur. The magnitude of h is determined by the net colloidal repulsion.
Assume, for example, that y = 5 mN m~^ and h= 10 nm. This results in AG^^^ =
5x10"^^ J or 125 kT, which would mean that the film cannot rupture. If h = 3 nm,
the result would be 11 kT. Assuming the film area to be 0.01 fim , this would lead,
according to [8.3.24], to a rupture probability of 10^^ s~^, i.e., immediate rupture.
However, the theory needs considerable modification. It has been pointed out, for
example, that the inter facial tension need not be constant during hole formation.
Moreover, the compression of the monolayers is counteracted by the surface dilational
modulus; a correction as given in [8.2.19] has to be applied.
Kabalnov-Wennerstrom theory^\ This theory applies for the regime where the
surfactant is an amphiphlle, We » 1, Ma > 1, and agitation is absent or weak.
When a hole is formed in a film at a small h value (fig. 8.26a), the local curvature of
the monolayer is very strong. This can go along with a considerable increase in bend-
ing Gibbs energy, which should be added to the de Vries value of AG^^^, especially if
the interfacial tension is low. Both the mean and the Gauss bending modulus are
important. The values of these moduli depend closely on the natural curvature of the
monolayer. The latter depends on the shape of the amphiphile molecules^^ A
(truncated) cone shape can lead to a high, and a cylindrical shape to zero natural cur-
vature. Moreover, a difference between film curvature and natural curvature leads to
an increase in interfacial tension.
All of these effects, as well as the Gibbs energy of compression of the monolayer,
are taken into account in the theory, which is therefore intricate. The theory is
especially successful for systems that exhibit a phase-inversion temperature (sees.
8.1b, sub (i) and 8.2c, sub (iv)), and nicely explains the mechanism causing phase-
inversion, as a function of surfactant properties.
VriJ-Scheludko theory^K As discussed in relation to [8.3.24], transverse thermal
waves can develop spontaneously on an interface, but waves of a much longer wave-
length are considered here. Such waves tend to be strongly damped, since they cause

The original article is, A. Kabalnov, H. Wennerstrom, Langmuir 12 (1996) 276. An updated
review is by A. Kabalnov, chapter 7 in B.P. Binks (1998), see General References.
See, e.g., J.N. Israelachvilli, Intermolecular and Surface Forces, 2""^ ed.. Academic Press
(1992).
^^ Developed independently by A. Vrij, Discuss. Faraday Soc. 42 (1966) 1966, and A.D.
Scheludko, Adv. Colloid Interface Sci. 1 (1967) 391.
EMULSIONS 8.87

local differences in the Laplace pressure. If, however, a net attractive force acts
between the two faces of a film, and if this force is stronger for a lower value of h, the
waves on the two surfaces can become coupled, as illustrated in fig. 8.26b. Van der
Waals forces acting across the film fulfil these criteria, provided that the film is thin
enough. The condition for coupling depends on a number of conditions; in most cases
of interest it is given approximately by

h<0.15p^^^ [8.3.27]
I r )
Assuming A = 5 x 10"^^ J , R = \ |im , and 7 = 3 mN m~^, the critical value of h is
about 5 nm. The condition for h^^ implies, inter alia, that the Vrij-Scheludko theory
only applies if We » 1.
Coupled waves can cause local thinning of the film; hence, the film can rupture at
an average h value that would provide stability, according to the theory discussed
above. Whether this occurs depends on a balance of forces. First, the Laplace pressure
gradients induced in the film cause flow of continuous phase from a thick to a thin
spot. In other words, the wave tends to be damped. The larger is the film radius R ,
the longer can be the wavelength of the film, and hence the smaller the curvature of the
interfaces for the same wave amplitude, the smaller the Laplace pressure gradient,
and the weaker the damping. Consequently, waves with a wavelength X ~ 2R will
dominate. Second, repulsive forces generally act between the faces of the film, and the
total interaction Gibbs energy as a function of film thickness, AG(h), must be taken
into account. The condition for the amplitude to grow, in which case the film will
eventually rupture (or form a metastable black film), is given by

^ < - ^ [8.3.28,
dh^ R^
Considering a DLVO type of interaction, [8.3.28] predicts that the film will always
rupture for a value of h corresponding to the primary minimum in the interaction
curve, whereas the film will generally be stable in the secondary minimum.
The theory given above has its limitations. First, it does not apply to very large
films, say R > 50 |im , as mentioned^^ This can be neglected for nearly all emulsions:
the drops are too small. Second, it is implicitly assumed that the interfaces are rigid.
In an extension of the theory^^ the effect of the surface dilational modulus is intro-
duced. It turns out that the K^ values encountered in practice are nearly always
sufficient to have a rigid interface. For the simple case, where the only colloidal
interaction working is van der Waals attraction, the condition is Kj >A/27ih^.

^^ See also, B.P. Radoev, A.D. Scheludko, and E.D. Manev, J. Colloid Interface Sci. 95 (1983)
254.
A. Vrij, F.T. Hesselink, J. Lucassen, and M. van den Tempel, Proc. Kon. Akad. Wetensch.
B73 (1970) 124.
8.88 EMULSIONS

Hence, the critical magnitude for the modulus is virtually always below 1 mN m~^ .
Only when the surfactant concentration is very low, may this value not be reached; if
so, coalescence can occur already at a large value of h , provided condition [8.3.27] is
fulfilled. Third, the theory would need modification if the main attractive force does
not increase with decreasing h , as is the case for depletion interaction.
Other aspects. The theories for film rupture do not generally provide good
quantitative prediction of coalescence rates, but most trends are predicted. However,
in some cases, variables other than those considered seem to control the coalescence
process. Examples follow.
A black film may form, rather than a hole in the film. Black films can have a
considerable life time. Whether this occurs in films between small emulsion droplets
and, if so, under what conditions, is poorly known. This is in contrast to isolated,
plane-parallel films, for which black film formation is an important stabilization
mechanism, see sec. 6.4d.
Some arnphiphiles can give rise to the formation of a laminar liquid crystalline
phase at a relatively low concentration; this also depends on the further composition
of the solvent. Such a surfactant may then form multilayers around emulsion drop-
lets, if its concentration is high enough, and such multilayers provide a very high
stability against coalescence^^ The explanation is still a matter of debate.
Polymers. The rupture theories discussed do not apply if the surfactant used is a
polymer. The author is not aware of a sound theory. Nevertheless, it is well known
that polymeric surfactants can provide long term stability (several years) against
coalescence, even for We »1 and in a sediment layer. This is presumably a result of
steric repulsion, acting over a relatively long range, keeping the droplets far away from
each other and thereby precluding film rupture. Other things being equal, polymers of
a higher degree of polymerization (in the range of, say, ten to several hundreds) tend
to provide greater stability.
Surfactant layer coherence. Several workers have observed that the coalescence
rate of emulsions is correlated with the value of the surface shear viscosity, 77^ (often
called by the misleading term, 'film strength'), as measured at a macroscopic
interfacial layer of the surfactant used in making the emulsion. An example is given by
water in crude oil emulsions in which the surfactant consists of high molar mass
asphaltenes in the mineral oil^^^ Another example is o/w emulsions stabilized by
various globular proteins'^^ some of these proteins are known to form intermolecular
cross links in an adsorption layer, which also appears to correlate with the value of
T/g . The correlations are not perfect, and some clear exceptions to the rule have been

^' See, e.g., S. Friberg, L. Mandell, and M. Larsson, J. Colloid Interface Set 29 (1969) 155.
^^ P. Becher, p. 257 in H.F. Eicke, G.D. Parfitt, Eds., Interfacial Phenomena in Apolar Media,
Marcel Dekker (1987).
^^ E. Dickinson, B.S. Murray, and G. Stainsby, J. Chem. Soc. Faraday Trans. I 84 (1988) 871.
EMULSIONS 8.89

observed. It has recently been reported^^ that failure, i.e., yielding or fracture, of the
adsorption layer can occur during the measurement of j]^, which implies that the
viscosity is measured in a 'destroyed' layer. In other words, the result is deceptive. It
may be better to directly measure two-dimensional stress and strain at failure and
correlate the results with proneness to coalescence. It may well be that strong
coherence of the adsorption layer prevents rupture of the film surfaces, and thereby
rupture of the film itself.

(iv) Concluding remarks. Several factors determine whether coalescence can occur,
and those dominating during emulsion formation and during storage are quite differ-
ent. During/ormation, the Marangoni number (at very short time scale) is the essen-
tial parameter. If Ma > 1, rapid drainage of the film between approaching droplets is
prevented, and coalescence will rarely occur. On the other hand, during storage the
Weber number is the essential parameter. If We < 1, the film formed between droplets
is generally too small in airea or not thin enough to allow film rupture.
There are a few exceptions to this dichotomy. If a 'finished' emulsion is agitated
during storage (at an intensity that is insufficient to cause droplet break-up) an
intermediate situation arises. This is, to some extent, discussed in subsec. (ii), but
only for simple-shear flow. For elongational and turbulent flows, the relations are
partly different. Another exception relates to emulsion formation and stability at
conditions close to th^ phase-inversion temperature. Here, We » 1, and any agitation
is quite weak; it follows that stability during formation and storage are, for the most
part, governed by the same variables. Finally, Pickering emulsions appear to provide
an exception; see sec. 8.4.
Some practical variables will now be considered. Unless mentioned otherwise, it is
assumed that the droplets are not too large, say < 10 )Lim, and are covered by a
monolayer of surfactant; agitation is assumed to be negligible.
Droplet size. For a larger value of d, the value of We is larger, and the area of the
film forming between two droplets will be larger, also at constant We. This will mean
faster coalescence, because the probability of film rupture is enhanced. This
phenomenon has been observed by several workers. One study concerned very dilute
protein stabilized emulsions, where the surface excess was much below its plateau
value, so that coalescence did occur. The oil drops were allowed to cream to an oil
layer on top of the emulsion. Drops reaching that position were observed by a
microscope, and the average time needed for coalescence of a drop with the oil layer
was noted. The time rsinged from about 35 s for 2 jam drops, to 2.5 s for 10 |im
drops^^ In another investigation, amphiphile-stabilized drops in a highly concentrated
emulsion (e.g., (p= 0.85, hence W e » 1 ) were studied. The coalescence rate was

^^ T. van Vliet, G.A. van Aken, M.A. Bos, and A.H. Martin, p. 176 in E. Dickinson, T. van Vliet,
Eds., Food Colloids, Biopolymers and Materials, Royal Soc. Chem. (2003).
^^ See E. Dickinson, et al. (1988), loc cit.
8.90 EMULSIONS

emulsion (e.g., (p= 0.85, hence W e » 1 ) were studied. The coalescence rate was
established by monitoring the average drop size, and the rate was found to be
proportional to d^ ^\
Furthermore, a larger drop diameter generally leads to faster sedimentation, and
hence to formation of a sediment (cream) layer, hence to increased We and a longer
time available for film thinning and rupture, hence to faster coalescence. Finally, for a
larger initial droplet size (and constant cp) fewer coalescence events are needed to
obtain a visible change.
Surfactant type. The surfactant type and concentration can, to a considerable
extent, determine the drop size distribution, which itself has a large effect on coales-
cence. Thus, it makes no sense to compare the effect of various surfactants on coales-
cence stability unless the comparison is carried out at the same droplet size (a rule
that is often broken, unfortunately).
The most important effect that the surfactant may have is to provide colloidal
repulsion. This affects the thickness of the film between drops, and hence the coales-
cence rate. It may also affect the rate of perikinetic or orthokinetic aggregation, and
thereby the coalescence rate: under some conditions (high We), drops cam coalesce
directly after they have aggregated. Naturally, solvent properties such as ionic strength,
pH, and solvent quality, can strongly affect the repulsive force and range.
Another essential effect is that the surfactant determines the interfacial tension.The
lower is the interfacial tension, the faster the coalescence tends to be: We is larger
and the theories of de Vries and Vrij-Scheludko predict a higher film rupture fre-
quency, as is indeed observed^\ The surfactant can also provide a high bending
modulus, which can counteract film rupture especially if the interfacial tension is quite
small.
Generally, surface active polymers provide better stability against coalescence than
do low M amphiphiles, since the range over which (steric) repulsion acts is larger,
and the interfacial tension is not very small, say, 10 mN m~^. It has often been
observed for protein stabilized emulsions that coalescence is induced by the addition
of some amphiphile that partly displaces protein from the interface, thereby lowering
the interfacial tension and, possibly, decreasing colloidal repulsion or layer coherence.
Finally, both amphiphiles and polymers, if present in sufficient concentration in the
continuous phase, can induce droplet aggregation by depletion interaction, with all its
consequences.
Volume fraction. If coalescence occurs, it will proceed faster with a higher volume
fraction of the disperse phase.

B. Deminiere et al., chapter 8 in B.P. Binks (1998), General References.


^^ A good example is given by, J.A.M.H. Hofman. H.N. Stein, J. Colloid Interface Set. 147
(1991)508.
EMULSIONS 8.91

8.4 Case Study: Pickering Emulsions ^^

Emulsions that are stabilized by adsorbed small solid particles have been mentioned
in sec. 8.3b, sub (ii), as they can provide considerable stability against Ostwald
ripening. Such adsorbed particles can also stabilize an emulsion against coalescence;
this mechanism is called Pickering stabilization . The particles should be located in
the o/w interface, protruding further into the continuous phase than in the disperse
phase. The latter condition implies that 0 < cir < 90°, where a is the contact angle as
measured in the continuous phase; see sec. 111.5.11c. Moreover, the Gibbs energy for
desorption of a particle needs to be large. For a sphere of radius a , the energy of its
transfer to the continuous phase is given by

AG^^o = ^^p Yo^a-cosaf [8.4.1]

Assuming y^^ = 0.03 N m~^, a= 60°, and a = 3 nm, the value of AG equals about
50 kT, which would mean that desorption due to Brownian motion is virtually
impossible. Hence, even quite small particles can be used. (For transfer of the particle
to the disperse phase, the minus sign in the equation must be replaced by a plus sign).
There is another factor that affects coalescence. If two drops are pushed close to-
gether, adsorbed particles can move laterally, away from the point of closest approach,
and this can cause formation of a bare patch on the drop(s), which may, in turn,
induce coalescence. The lateral motion can be counteracted by the formation of
surface tension gradients, which needs a significant surface dilational modulus. The
value of K^ depends on, (i), the particle surface coverage, which must be large, say,
6>0.7 (depending on particle shape and on lateral interaction forces between the
adsorbed particles); and, (ii), on the presence, type, and surface excess of surfactant(s)
in the interface.
However, the stability and the formation of the emulsion cannot be seen as
independent processes. To make a Pickering emulsion, agitation is needed, but the
intensity of agitation mostly does not determine drop size. The limiting factors are the
concentration and size of the solid particles. The droplet diameter always exceeds the
particle diameter, by a factor ranging mostly between 30 and 300, even if the particle
concentration is not limiting. It has frequently been observed that the contact angle
should not be much below 90° to obtain relatively small droplets ( d ~ 5 0 a ). The
explanation of these phenomena is yet unclear, to the author's knowledge.
In most of the older studies on Pickering stabilization, relatively large solid par-

^^For the most part, based on S. Levine, B.D. Bower, and S.J. Partridge, Colloids Surfaces 38
(1989) 325 and 345; D.E. Tambe, M.M. Sharma, Adv. Colloid Interface Sci. 52 (1994) 1; B.P.
Binks, S.O. Lumsdon, Langmuir 16 (2000) 8622; and N.X. Yan, M.R. Gray, and J.H. Masliyah,
Colloids Surfaces A193 (2001) 97.
2'After S.U. Pickering, J. Chem. Soc. 91 (1907) 2001, although W. Ramsden, Proc. Roy. Soc.
(London) 72 (1903) 156, was earlier in describing the phenomena.
8.92 EMULSIONS

tides were used, diameter 0(1 [im ), resulting in large drops. Generally, a surfactant is
added, ostensibly to change and fine tune the contact angle. It cannot be ruled out,
however, that the surfactant also helps in forming y gradients during emulsification,
thereby counteracting drop re-coalescence; see sec. 8.2c, sub (vi). More recently,
several far smaller particles (about 10 nm or even less) have become available, and
some of these nanoparticles, especially those made of silica, can be tailor made so that
they produce the desired contact angle. Pickering emulsions can now be made in the
absence of an added surfactant. Probably, the nanoparticles in the interface can prod-
uce a significant surface pressure, owing to their relatively large molar surface excess.
For particles of 5 nm diameter and 0= 0.75, [1II.3.4.39] yields TT = 2.0 mN m ' ^
Moreover, the particles are large enough to form a true Langmuir monolayer. Hence, a
small but significant surface dilational modulus can result, allowing the formation of
interfacial tension gradients, provided that the droplets are small.
Another difference is that the coarse emulsions tend to be unstable, unless weak
attractive forces act between the particles in the interface. The latter has been shown
to result in larger values of the surface dilational modulus. The attraction should not
be strong, since the particles would then aggregate in solution, thereby precluding the
formation of an emulsion, since the aggregates are generally too large to allow full
coverage of the drops. Hence, careful optimization of product properties and process
conditions are needed to obtain a stable emulsion. Such problems do not (or need
not) occur when nanoparticles are used. This may, again, be due to the nanoparticles
producing a sufficiently large surface dilational modulus. Moreover, the drops
obtained are far smaller (provided that the agitation is sufficiently intense) and hence
the Weber number for coalescence will be smaller. For nanoparticles, the contact angle
should also be close to 90° to allow formation of small drops.
Theories have been proposed to explain some of the phenomena observed, and
these have certainly enhanced understanding. However, a definitive and all embracing
theory of the formation and stability of Pickering emulsions is not yet available. This
means that the 'state of the art' is , as for nearly all types of emulsions, that the applic-
ation of the fundamentals of surface and colloid science has produced considerable
and useful understanding of the processes involved, and of the properties resulting,
but several significant topics still need further research.

8.5 General References

R. Aveyard, B.P. Binks, and J.H. Clint, Emulsions, Stabilized Solely by Colloidal
Particles. Adv. Colloid Interface Set 100 (2003) 503-546. (Review, emphasizing the
wetting and thermodynamic properties of well-defined particles at the o/w interface;
the paper also includes a section on triple emulsions (w-o-w and o-w-o).)
EMULSIONS 8.93

P. Becher, Emulsions: Theory and Practice, Reinhold Publ. (1953). (Presumably


the first book giving substantial understanding and a range of properties.)

B.P. Sinks, Ed., Modern Aspects of Emulsion Science, Royal Soc. Chem. (1998).
(Most aspects discussed in this chapter are treated, largely based on fundamental
theory aind in great detail. Especially recommended are chs. 1, by B.P. Binks, on
recent advances; 2, by P. Walstra, P.E.A. Smulders, on emulsion formation; 5, by E.
Dickinson, on emulsion rheology; 6, by B.W. Brooks et al., on phase-inversion; 7, by
A.S. Kabalnov on coalescence; 8, by B. Deminiere et al., on coalescence in concen-
trated emulsions; 9, by J.G. Weers, on Ostwald ripening and related phenomena; and
10, by D.M. Petsev, on the various interaction forces between droplets.)

A.K. Chesters, The modelling oj coalescence processes in fluid liquid dispersions:


A review of current understanding. Trans. Inst. Chem. Eng. 69-A (1991) 259. (Avery
clear and complete review.)

J.T. Davies, Turbulence Phenomena, Academic Press (1972). (Much information


on droplet break-up, particularly in the turbulent regime.)

Encyclopedia of Emulsion Technology, Vol. 1, Basic Theory (1983); Vol. 2, Applic-


ations (1985); Vol. 3, Basic Theory Measurement Applications (1988); Vol. 4 (1996);
P. Becher, Ed., Marcel Dekker. (Contains several chapters on themes discussed in this
chapter and also treats a range of apphcations.)

LB. Ivanov, P.A. Kralchevsky, Stability of emulsions under equilibrium and


dynamic conditions, Colloids Surfaces A, 128 (1997) 155. (A critical analysis of
theories and results on hydrodynamic and thermodynamic aspects.)

LB. Ivanov, K.D. Danov, and P.A. Kralchevsky, Flocculation and coalescence of
micron sized emulsion droplets. Colloids Surfaces A152 (1999) 161. (An extension of
part of the previous article.)

V.G. Levich, Physicochemical Hydrodynamics, Prentice-Hall (1962). (Includes


much information on droplet comminution under various conditions.)

E.H. Lucassen-Reynders, Ed., Anionic Surfactants: Physical Chemistry of Surf act-


ant Action, Surfactant Series 1 1 , Dekker Marcel (1981). (Contains treatments of
fundamental aspects important for emulsion formation and stability.)

P. Sherman, Ed., Emulsion Science, Academic Press, London (1968). (Gives funda-
mentals of emulsion formation, stability, and various properties. Especially chapter 1,
by E.S.R. Gopal, on emulsion formation is still worth reading.)
8.94 EMULSIONS

H.N. Stein, Ed., The Preparation of Dispersions, Chem. Eng. Sci. 46 (2), (1993),
pp. 201-460. (A symposium report containing seven contributions on emulsion
formation.)

P. Walstra, Formation of Emulsions, ch. 2, pp. 57-127, in; P. Becher, Ed., Ency-
clopedia of Emulsion Technology, Vol. 1, Marcel Dekker (1983). (A fairly compre-
hensive review.)

P. Walstra, Principles of Emulsion Formation. Chem. Eng. Sci. 4 6 (1993) 333-350.


(A review with some emphasis on the time scales of the various processes.)

P. Walstra, Emulsion Stability, ch. 1, pp. 1 62, in P. Becher, Ed., Encyclopedia of


Emulsion Technology, Vol. 4, Marcel Dekker, 1996. (A review)

You might also like